Download as pdf or txt
Download as pdf or txt
You are on page 1of 212

Managing Forest Ecosystems

Víctor Resco de Dios

Plant-Fire
Interactions
Applying Ecophysiology to Wildfire
Management
Managing Forest Ecosystems

Volume 36

Series Editors
Klaus von Gadow, Georg-August-University, Göttingen, Germany
Timo Pukkala, University of Joensuu, Joensuu, Finland
Margarida Tomé, Instituto Superior de Agronomía, Lisboa, Portugal
Well-managed forests and woodlands are a renewable resource, producing essential
raw material with minimum waste and energy use. Rich in habitat and species
diversity, forests may contribute to increased ecosystem stability. They can absorb
the effects of unwanted deposition and other disturbances and protect neighbouring
ecosystems by maintaining stable nutrient and energy cycles and by preventing soil
degradation and erosion. They provide much-needed recreation and their continued
existence contributes to stabilizing rural communities.
Forests are managed for timber production and species, habitat and process
conservation. A subtle shift from multiple-use management to ecosystems
management is being observed and the new ecological perspective of multi-­
functional forest management is based on the principles of ecosystem diversity,
stability and elasticity, and the dynamic equilibrium of primary and secondary
production.
Making full use of new technology is one of the challenges facing forest
management today. Resource information must be obtained with a limited budget.
This requires better timing of resource assessment activities and improved use of
multiple data sources. Sound ecosystems management, like any other management
activity, relies on effective forecasting and operational control.
The aim of the book series Managing Forest Ecosystems is to present state-of-
the-art research results relating to the practice of forest management. Contributions
are solicited from prominent authors. Each reference book, monograph or
proceedings volume will be focused to deal with a specific context. Typical issues
of the series are: resource assessment techniques, evaluating sustainability for even-
aged and uneven-aged forests, multi-objective management, predicting forest
development, optimizing forest management, biodiversity management and
monitoring, risk assessment and economic analysis.

More information about this series at http://www.springer.com/series/6247


Víctor Resco de Dios

Plant-Fire Interactions
Applying Ecophysiology to Wildfire
Management
Víctor Resco de Dios
School of Life Science and Engineering
Southwest University of Science and Technology
Mianyang, China
Crop and Forest Sciences and JRU CTFC-AGROTECNIO
Universitat de Lleida
Lleida, Spain

ISSN 1568-1319     ISSN 2352-3956 (electronic)


Managing Forest Ecosystems
ISBN 978-3-030-41191-6    ISBN 978-3-030-41192-3 (eBook)
https://doi.org/10.1007/978-3-030-41192-3

© Springer Nature Switzerland AG 2020


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of
the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilms or in any other physical way, and transmission or information
storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology
now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the
editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Switzerland AG.
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
In memory of my late grandmother, Concha
Vidal Cerdeiros (1930–2019), who did not
live enough to see this book, and to my boys,
Hernán and Martín, and girl, María, who
had to undergo the entire process.
Acknowledgments

I do not have a control over my research interests anymore. On the contrary, I am a


victim of my research, which is the one telling me where I should be going to next.
This all started after Dave Williams, from the University of Wyoming, introduced
me to the world of stress physiology and after José M Moreno, from the University
of Castilla-la Mancha, introduced me to the world of fire. I joined the faculty of
Western Sydney University and shared an office (briefly) with Pyrogeographer
Matthias Boer, and here is when everything ignited. Matthias introduced me to Ross
Bradstock from the University of Wollongong, and the three of us started to explore
together the linkages between physiological ecology and wildfire science. During
that time, I was extremely fortunate in working with Rachael Nolan, Luke Collins,
and Gabriele Caccamo, who taught me further on plant-fire interactions. In more
recent years, I have benefited from the interactions with my colleagues at the
University of Lleida and, in particular, with Jordi Voltas and Marc Castellnou. This
book is just the result of my learning from working with these colleagues. If there is
something meritorious in this book, it is thanks to them. I can only take credit for
the mistakes.
I wish to acknowledge the colleagues who have generously donated their time to
review the chapters in this book. I thank Rachael Nolan, David Tissue, and Matthias
Boer from Western Sydney University, Marta Yebra from Australia National
University, Jordi Voltas from the University of Lleida, José Manuel Moreno from
the University of Castilla-la Mancha, César Morales-Molino from the University of
Bern, Javier Madrigal from the National Institute for Agricultural Research, and
Juan Picos from the University of Vigo. I also thank Jaime Coello for his advice on
merging pyrophysiology with wildland management, Lluís Coll for his discussions
on post-fire dynamics, and José Luis Ordoñez, José Carlos López Sánchez, Javier
Madrigal, and Rachael Nolan for kindly sharing one or more figures.
I also wish to thank Takeesha Moerland-Torpey, Valeria Rinaudo, Malini
Arumugam, and the Springer editorial team for their encouragement to write this
and their help and patience throughout the process.
This book seeks to merge two different disciplines, plant physiological ecology
and wildfire science. Works that seek to join different disciplines often need to make

vii
viii Acknowledgments

compromises to merge the differences in standards and criteria across fields. This
often leads to criticisms from both sides, and I do not expect this book to be an
exception. I would thus like to end by thanking you for reading and, hopefully, criti-
cizing this work so we can jointly develop the emerging field of pyrophysiology to
help solve the wildfire problem, one of the major environmental challenges of
our days.
This book presents an original contribution except for some pictures, as noted,
and for Chap. 8, which is largely a republication of a recently published journal
article led by the book author (Karavani et al. 2018), and it has only been slightly
updated to keep it on track with more recent developments in the field.
Contents

1 Introduction����������������������������������������������������������������������������������������������    1
1.1 Pyrophysiology: The Ecophysiology of Wildfires����������������������������    1
1.2 Pyrophysiology and the Drivers of Fire Activity������������������������������    3
1.3 Pyrophysiology and Pyrogeography ������������������������������������������������    6
1.4 Fire in Ecosystems����������������������������������������������������������������������������    7
1.5 Fire in Societies��������������������������������������������������������������������������������    9
1.6 Scope and Structure of This Book����������������������������������������������������   10
References��������������������������������������������������������������������������������������������������   11
2 Fire Regimes Across Space����������������������������������������������������������������������   15
2.1 The Fire Regime ������������������������������������������������������������������������������   15
2.2 Fire Type by Fuel Layer Carrying the Fire ��������������������������������������   16
2.2.1 Surface Fires ������������������������������������������������������������������������   16
2.2.2 Crown Fires��������������������������������������������������������������������������   17
2.2.3 Ground Fires ������������������������������������������������������������������������   18
2.3 Intensity and Severity������������������������������������������������������������������������   18
2.3.1 Fire Intensity ������������������������������������������������������������������������   18
2.3.2 Fire Severity��������������������������������������������������������������������������   21
2.4 Temporal Patterns ����������������������������������������������������������������������������   22
2.5 Seasonality����������������������������������������������������������������������������������������   24
2.6 Mosaic����������������������������������������������������������������������������������������������   25
2.7 Pyromes��������������������������������������������������������������������������������������������   26
References��������������������������������������������������������������������������������������������������   27
3 Fire as an Earth System Process������������������������������������������������������������   31
3.1 The Earth as a System and the Role of Fire��������������������������������������   31
3.1.1 Fire and Oxygen��������������������������������������������������������������������   32
3.1.2 Fire-Climate Feedbacks��������������������������������������������������������   34
3.2 Fire Before Humans��������������������������������������������������������������������������   38
3.2.1 Early Fires in the Paleozoic��������������������������������������������������   38
3.2.2 Wildfires in the Mesozoic ����������������������������������������������������   40

ix
x Contents

3.3 Fire in the Quaternary ����������������������������������������������������������������������   41


3.3.1 Humans Increased Burned Area in the Pleistocene
and Decreased It During the Holocene in Africa������������������   41
3.3.2 Human Alterations to the Fire Regime During
the Holocene ������������������������������������������������������������������������   42
3.4 Fire in the Anthropocene������������������������������������������������������������������   45
3.4.1 Changes in the Fire Regime from Recent
Human Activity��������������������������������������������������������������������   45
3.4.2 Wildfires and Health ������������������������������������������������������������   46
References��������������������������������������������������������������������������������������������������   47
4 The Evolution of Physiological Adaptations
in a Flammable Planet ����������������������������������������������������������������������������   53
4.1 Introduction��������������������������������������������������������������������������������������   53
4.2 Fire as a Selective and Mutagenic Agent������������������������������������������   56
4.3 Fire and the Evolution of Strategies Enhancing Recruitment����������   58
4.3.1 Serotiny ��������������������������������������������������������������������������������   58
4.3.2 Seed Germination�����������������������������������������������������������������   60
4.4 Fire and the Evolution of Strategies Enhancing Persistence������������   62
4.4.1 Resprouting ��������������������������������������������������������������������������   62
4.4.2 Meristem Protection��������������������������������������������������������������   64
4.5 Fire and the Evolution of (Anti-)flammability����������������������������������   66
4.6 Adaptations to Fire and Physiological Trade-Offs����������������������������   68
References��������������������������������������������������������������������������������������������������   70
5 Environmental Plant Responses and Wildland Fire Danger ��������������   75
5.1 Introduction��������������������������������������������������������������������������������������   75
5.2 Fuel Moisture as Limiting Fire Spread ��������������������������������������������   76
5.3 Wildfire Danger Indices Provide Limited Information
on Environmental Fire Danger����������������������������������������������������������   77
5.4 Drivers of Dead Fuel Moisture ��������������������������������������������������������   78
5.5 Drivers of Live Fuel Moisture����������������������������������������������������������   79
5.6 Problems in Measuring and Estimating Fuel Moisture
in Plant Canopies������������������������������������������������������������������������������   83
5.7 Fuel Moisture and Fire Occurrence��������������������������������������������������   85
5.8 Relationships Between Drought and Fire Occurrence����������������������   87
References��������������������������������������������������������������������������������������������������   88
6 Plant Carbon Economies and the Dynamics of Wildland Fuels����������   93
6.1 Wildland Fuels and Fire��������������������������������������������������������������������   93
6.2 Development of Live Wildland Fuels ����������������������������������������������   94
6.2.1 Different Life-Forms������������������������������������������������������������   94
6.2.2 Variation in Growth and Allocation Patterns Is
Coordinated with Post-fire Regeneration Strategy���������������   97
6.2.3 Plant Development, Canopy Architecture,
and Competition as Drivers of Crown Fuel Disposition������   99
Contents xi

6.3 Seasonal Changes in the Flammability of Live Fuels���������������������� 103


6.3.1 Biogenic Volatile Organic Compounds Have
an Uncertain Effect over Fire Behavior�������������������������������� 103
6.3.2 Changes in Fuel Inorganic Chemistry and Leaf
Anatomy That Potentially Affect Fire Behavior ������������������ 104
6.4 Development of Dead Wildland Fuels���������������������������������������������� 107
6.4.1 Leaf Senescence�������������������������������������������������������������������� 107
6.4.2 Self-Pruning, Self-Thinning, Bark Shedding,
and Plant Death �������������������������������������������������������������������� 109
6.4.3 Deposition and Decomposition�������������������������������������������� 110
References�������������������������������������������������������������������������������������������������� 111
7 Effects of Fire on Plant Performance ���������������������������������������������������� 117
7.1 Introduction�������������������������������������������������������������������������������������� 117
7.2 Direct Effects of Fire on Plant Performance������������������������������������ 119
7.2.1 Crown Defoliation���������������������������������������������������������������� 120
7.2.2 Phloem and Cambium Charring������������������������������������������� 121
7.2.3 Fire-Induced Cavitation and Sap Transport
Impairment���������������������������������������������������������������������������� 122
7.2.4 Root Consumption���������������������������������������������������������������� 123
7.3 Resprouting �������������������������������������������������������������������������������������� 123
7.3.1 Location and Protection of the Bud Bank���������������������������� 124
7.3.2 Resources Limiting Resprouting������������������������������������������ 125
7.4 Indirect Effects of Fire on Plant Performance���������������������������������� 127
7.4.1 Growth and Survival ������������������������������������������������������������ 127
7.4.2 Reproduction������������������������������������������������������������������������ 128
References�������������������������������������������������������������������������������������������������� 129
8 Forest Succession, Alternative States, and Fire-Vegetation
Feedbacks�������������������������������������������������������������������������������������������������� 133
8.1 Post-fire Changes in Community Dynamics������������������������������������ 133
8.2 Resilience to Fire at the Population Level���������������������������������������� 135
8.2.1 Propagule Availability���������������������������������������������������������� 136
8.2.2 Propagule Establishment������������������������������������������������������ 141
8.2.3 Propagule Survival���������������������������������������������������������������� 142
8.2.4 Mycorrhizal Networks���������������������������������������������������������� 145
8.3 State Transitions After Fire and Changes in Community
Dominance���������������������������������������������������������������������������������������� 146
8.3.1 Short-Term Transitions �������������������������������������������������������� 146
8.3.2 Long-Term Dynamics ���������������������������������������������������������� 147
References�������������������������������������������������������������������������������������������������� 149
9 Pyrophysiology and Wildfire Management ������������������������������������������ 155
9.1 Objectives of Wildfire Management ������������������������������������������������ 155
9.2 Fire Generations�������������������������������������������������������������������������������� 156
9.3 Fire Spread Patterns and Preventive Measures �������������������������������� 159
xii Contents

9.4 Pyrophysiology and Wildfire Prevention������������������������������������������ 161


9.4.1 Development of Economically Viable Landscapes
Through Ecophysiology�������������������������������������������������������� 162
9.4.2 Silvicultural Treatments Diminishing Water Stress�������������� 162
9.4.3 Fuel Reduction Treatments �������������������������������������������������� 164
9.5 Green Firebreaks������������������������������������������������������������������������������ 166
9.6 Pyrophysiology: Plant Flammability and Fire Behavior������������������ 166
9.7 Pyrophysiology and Fire Behavior Modeling���������������������������������� 169
9.8 Pyrophysiology and Post-fire Management and Restoration������������ 170
References�������������������������������������������������������������������������������������������������� 172
10 Global Change, Pyrophysiology, and Wildfires������������������������������������ 177
10.1 Fire and Climate Change���������������������������������������������������������������� 177
10.1.1 Fire in Tropical Forests Is Linked to Land Cover
and Land Use���������������������������������������������������������������������� 179
10.1.2 Woody Thickening Affects Tropical and Subtropical
Savanna Fires���������������������������������������������������������������������� 180
10.1.3 Unimodal Evolution of Wildfires
in Mediterranean Biomes���������������������������������������������������� 183
10.1.4 Increases in Forest Fire in Temperate
and Boreal Forests�������������������������������������������������������������� 184
10.2 Future Wildfires, National Security, and Civil Protection�������������� 186
10.2.1 Urbanization and Wildfires ������������������������������������������������ 186
10.2.2 The Challenge of Mega-Fires, Increasing Fire
Intensity and Pyroconvection���������������������������������������������� 187
10.3 Wildfires and Global Change���������������������������������������������������������� 188
10.3.1 Wildfires and Plant Invasions���������������������������������������������� 188
10.3.2 Wildfires and Biodiversity�������������������������������������������������� 190
References�������������������������������������������������������������������������������������������������� 193

Index������������������������������������������������������������������������������������������������������������������ 199
Chapter 1
Introduction

Abstract The physiological ecology of plant-fire relations, or pyrophysiology,


seeks to provide a mechanistic and predictive understanding of the reciprocal inter-
actions that exist between plants and fire. That is, by understanding the processes
underlying plant structure and function, we may gain further insight on how wild-
fires affect plants as well as on how the physiological traits that affect wildfires
develop. This chapter presents a general introduction to the topics of pyrophysiol-
ogy and to wildfire science, and it ends by presenting the book structure.

1.1 Pyrophysiology: The Ecophysiology of Wildfires

Wildfires annually burn 350 Mha, or 2.3%, of global land area (Fig. 1.1, Table 1.1)
(Chuvieco et al. 2016; Giglio et al. 2013). Wildfires have played a major role in
shaping the structure and function of terrestrial ecosystems throughout Earth’s his-
tory. Vascular plants first appeared at the transition between the Silurian and the
Devonian, about 420 million years ago, and wildfires were one of the first and
unavoidable consequences of vascular plant expansion (Glasspool et al. 2006).
Wildfires appeared on land much before humans did, and they will likely survive us.
It makes little sense to conceive wildfires as an agent to be eradicated. On the con-
trary, wildfires are one of our main allies to achieve a healthy environment.
Plants make the fuel for wildfires and fuel management is the primary tool for
fire management. Wildland fuels have traditionally been approached from two dif-
ferent perspectives. One is the applied perspective that forestry, or fire behavior,
provides, and the other is a more basic view based on plant ecology. Attempts at
linking both perspectives are still rare (Keane 2015), and an effort towards under-
standing the mechanisms governing wildland fuel dynamics is missing.
Here we will discuss the physiological ecology of plant-fire relations, or pyro-
physiology. That is, we will explore to what extent may an understanding of plant
structure and function help bridge the gap between wildland fuel management, fire
behavior, and plant ecology.
Plant physiological ecology and wildfire science have both a long tradition of
research, but they have mostly developed in parallel, and only in a few occasions

© Springer Nature Switzerland AG 2020 1


V. Resco de Dios, Plant-Fire Interactions, Managing Forest Ecosystems 36,
https://doi.org/10.1007/978-3-030-41192-3_1
2 1 Introduction

Fig. 1.1 Observed (1995–2016) mean annual fractional burned area. Reproduced from Boer
et al. (2019)

Table 1.1 Annual burned area statistics for different world regions
Land area Burned area % land area annually % global
Region (Mha) (Mha) burned burned area
Boreal N America 1120 2.2 0.20 0.63
Temperate N America 780 1.8 0.23 0.51
Central America 270 1.8 0.67 0.51
N Hemisphere South 300 2.6 0.87 0.74
America
S Hemisphere South 1470 18.7 1.27 5.35
America
Europe 700 0.7 0.10 0.20
Middle East 1190 0.8 0.07 0.23
Northern Hemisphere 1470 117.7 8.01 33.66
Africa
Southern Hemisphere 980 125 12.76 35.74
Africa
Boreal Asia 1520 5.6 0.37 1.60
Central Asia 1800 13.6 0.76 3.89
Southeast Asia 660 7 1.06 2.00
Equatorial Asia 270 1.6 0.59 0.46
Australia and New 790 50.2 6.35 14.36
Zealand
Data from Giglio et al. (2013) and van der Werf et al. (2017)

have both disciplines converged. There are some notable exceptions to this, such as
attempts at explaining variation in fuel moisture from ecophysiological principles
(Nelson 2001) or studies on post-fire plant survival and growth (Dickinson and
Johnson 2001; Knapp 1985), resprouting (Pate et al. 1990), and regeneration
(Thanos and Georghiou 1988), to name a few. The number of studies on the eco-
physiology of wildland fuels addressing both how physiological fuel traits affect
1.2 Pyrophysiology and the Drivers of Fire Activity 3

fire and how fire affects plant physiology has increased notably in recent years
(Jolly and Johnson 2018; Bar et al. 2019; Nolan et al. 2018; West et al. 2016; Bond-
Lamberty et al. 2007; Parra and Moreno 2017; Hood et al. 2018; Kavanagh et al.
2010; Niccoli et al. 2019). A comprehensive synthesis on the physiological princi-
ples jointly governing plant growth, fuel development, and dynamics and how they
feedback to affect wildfire activity has been missing.
Pyrophysiology seeks to understand the reciprocal interactions between plants
and wildfires, and it may be viewed as a subdiscipline both of plant physiological
ecology and also of wildfire science. From a physiological perspective, pyrophysi-
ology studies the physiological mechanisms underlying the effects of wildfires on
plants. How plants respond to different fire regimes: how they regenerate after fire
and grow and protect themselves from fire would be one of the main concerns of
pyrophysiology. From the perspective of wildfire science, pyrophysiology seeks to
understand how plants affect fire: understanding the development of fuel traits that
affect fire danger, behavior, and fire regime overall is also one of the main goals of
pyrophysiology.
Just like with any other discipline, it is important to set the boundaries of pyro-
physiology. Physiology is a reductionist discipline. It seeks to understand macro-
scopic patterns based upon properties at lower scales. Wildfires arise from
interactions between the biosphere and the atmosphere. Pyrophysiology is mostly
concerned with understanding fuel properties: how they are affected by, and how
they affect, wildfires. Understanding other aspects of wildfire science, such as fire
behavior to name an example, requires integration with other disciplines, including
meteorology and fluid mechanics. Here we will explore the new knowledge that can
be gained by jointly studying wildfires and ecophysiology, but we will also discuss
what aspects of the fire regime cannot be understood from a purely ecological or
physiological perspective.
This initial chapter will serve the joint purposes of introducing the reader to
wildfires and also of introducing the structure of this book and the different aspects
we will cover.

1.2 Pyrophysiology and the Drivers of Fire Activity

To understand the scope of pyrophysiology, we must understand the drivers of fire


activity. At a first approximation, landscape fires may be viewed as dependent on
four conditions (Fig. 1.2). These conditions, or thresholds, may be viewed as
switches that are connected in series: fires will only occur once they are all switched
on (Bradstock 2010). The first switch is that of biomass accumulation, as fires can
only propagate when there is enough biomass to be burned (Bradstock 2010).
Within the context of wildfires, biomass is referred to as fuel. The set of fuel char-
acteristics that affect fire activity is known as the fuel complex (see Chap. 6). In a
nutshell, wildfires require the accumulation of large fuel loads, showing spatial con-
tinuity across landscapes. Many forests and shrublands, particularly those lacking
4 1 Introduction

(Decades to years)
Long timescales
Fuel No
build-up No fire
(1)

Yes

Fuel dry- No
down No fire
Medium timescales
(Weeks to days)

(2)

Yes

Critically
flammable
landscape

No
Ignition
No fire
(3)

Yes
Short timescales
(Days to hours)

Fire start

No
Fire Environment Minor fire run
(4)

Yes

Major
landscape fire

Fig. 1.2 Fire activity depends upon four switches acting in series. Pyrophysiology is particularly
important as a driver of the first two. Modified from Boer et al. (2017)

management, show large accumulations of fuel that show potential for carrying
large fires.
However, the existence of large accumulations of fuel per se does not suffice to
support large wildfires as the fuel may be too wet to be available for fire. The second
1.2 Pyrophysiology and the Drivers of Fire Activity 5

switch is thus fuel availability (Bradstock 2010) which, in this book, will be defined
as the moisture status of the fuel complex (see Chap. 5). Landscape-scale flamma-
bility, or the capacity of landscapes to burn, depends upon the interaction between
fuel load and fuel availability. Critical flammability thresholds are thus reached
when landscapes with large fuel loads are dry enough to sustain large wildfires
(Boer et al. 2017). However, and perhaps counterintuitively, it is not the availability
of live fuel (i.e., plants) that will determine landscape flammability but the status of
dead fuel (i.e., litter) that is often more important. In fact, the structure and avail-
ability of the very fine (less than 6 mm thick) and fine (6–25 mm) dead fuel particles
often determine the initial patterns of fire spread and landscape flammability. This
explains partly why large wildfires often occur under hot days: the moisture of dead
fine and very fine fuels responds very rapidly to variations in vapor pressure deficit
(Resco de Dios et al. 2015), a function of temperature and relative humidity (see
Chap. 5). However, dry plant canopies often set the stage for extremely large fires
(Nolan et al. 2016).
The third switch is an ignition source. As Bruce Springsteen’s Dancing in the
Dark says, “you can’t have a fire without a spark.” However, and this distinction is
important, a spark is not a fire. We often hear in the news statements like “the wild-
fire was caused by arson,” but the arsonist is only responsible for the ignition, not
for the wildfire. The ignition is just one of the four switches that triggers wildfire
activity, and they are only successful in becoming wildfires when the other three are
also on. Statements like “fires are caused by arson” are, at least sometimes, crude
attempts to oversimplify the complex phenomenon of wildfires by searching for a
scapegoat. At any rate, the vast majority of ignitions often have a human origin.
Consequently, ignitions are intrinsically difficult to predict, beyond some basic heu-
ristic rules (including distance to roads and to populated areas and land use), and
show a strong random component (Costafreda-Aumedes et al. 2017).
In landscapes that have crossed the critical flammability threshold (switches 1
and 2), fires will start after an ignition source (switch 3), but whether or not they will
become major landscape fires will then depend upon the fire environment (switch 4)
(Bradstock 2010). The fire environment constitutes the series of abiotic factors that
contribute to fire spread and propagation and is driven by the combination of meteo-
rological conditions at the time of the fire (fire weather) along with topographic
features (slope, aspect, relief, etc.). High temperatures and low humidity aid with
ignition and flame transport and wind contributes to fire spread by enhancing ambers
and by directly affecting flame length and depth (Bradstock 2010). Topographic
features may further enhance fire propagation as steep slopes favor fire propagation
through convection, radiation, and conduction. Additionally, complex terrain favors
the existence of topographic winds that may create locally unstable meteorological
conditions, partly decoupled from large-scale synoptic weather patterns.
An important feature of the four switches concept is the different temporal
domains acting on each switch. Fuel load results from cumulative vegetation growth
and varies at yearly to decadal scales. Fuel availability may change seasonally, often
within days or weeks, and transitions from non-flammable to flammable landscapes
6 1 Introduction

are relatively quick. Ignitions and fire weather often change at very short timescales
and represent the most variable component of fire activity (Boer et al. 2017).
Pyrophysiology is focused on explaining variations in switches 1 and 2: it is only
through physiological ecology that one may understand and develop mechanistic
predictions on changes in fuel load and fuel availability. And this will be the main
focus of the book.

1.3 Pyrophysiology and Pyrogeography

A corollary from the four switches concept is that potential fire activity across land-
scapes will vary as a function of interactions between flammability (fuel load and
availability), productivity, and aridity (Fig. 1.3). The “intermediate fire-­productivity”
hypothesis (Pausas and Bradstock 2007; Pausas and Ribeiro 2013; Keeley et al.
2012) proposes that there is a trade-off between productivity and aridity and shows
how relationships between fire and productivity, and between fire and aridity, are
hump-shaped (Fig. 1.3). Peaks in global fire activity occur at intermediate levels of
productivity and aridity, while minimum fire activity is observed where (i) produc-
tivity is low and aridity is high (deserts), as there is not enough biomass (i.e., fuel)
to carry a fire, or (ii) productivity is high and aridity is low (tropical rainforests), as
fuels are seldom dry enough to burn.
This may be seen in Table 1.1, which indicates global variation in burned area
across regions. Almost 70% of global burned area concentrates in African savannas,
where 10% of the territory burns annually. Australia follows as the second most
flammable continent, accounting for 14% of global burned area, and where 6% of
its territory burns annually. The percent of land area annually burned for the other
world regions is at or below 1%. The physiology of plants in areas burning at high
frequency shows adaptations for persistence, such as resprouting, whereas plants in
areas with smaller fire frequency additionally show adaptations for recruitment,
such as seed banks. The regeneration strategy is linked to many aspects of the envi-
ronmental and ecological physiology of plants, and this will be discussed in Chaps.
7 and 8.
Human populations often tend to decrease fire activity because putting off all
fires as soon as they appear is the most common response to fire activity (fire sup-
pression) (Andela et al. 2017). Consequently, high human pressure over the fire
regime in the Mediterranean and other areas may have decreased fire activity beyond
its potential, creating a “fire deficit” as a larger proportion of the land would nor-
mally burn under a more natural fire regime (Marlon et al. 2012). The natural drivers
of the spatial variation in fire activity are interactions between productivity and arid-
ity and their effect on fuel load and availability. However, human activity has dra-
matically modified patterns of fire activity to the point where human aspects are as
important as (if not more than) the natural drivers. That is, humans have altered land
use and land management, along with ignition and spread patterns, to the point
1.4 Fire in Ecosystems 7

Fig. 1.3 Fire activity


varies across productivity
and aridity gradients

Fire activity
following a unimodal
pattern. Modified from
Pausas and
Bradstock (2007)

ARIDITY
PRODUCTIVITY
Deserts TREE COVER Forests

where the probabilities of the four switches being concomitantly on have been sig-
nificantly altered. Human-fire interactions will be covered in Chaps. 9 and 10.
Wildfires are one of the major de facto management actors in terrestrial ecosys-
tems. Particularly in the developed part of the world, or in areas where us humans
have forgotten about our wildlands and forests, wildfires will act as one of the major
drivers of landscape and physiognomic changes. Catastrophic wildfires will always
be there. However, an increasing frequency is the result of accumulated decades of
poor land management in Mediterranean ecosystems.

1.4 Fire in Ecosystems

At a first approximation, wildfires act as major “herbivores” as they consume a large


portion of the biomass available (Bond and Keeley 2005). That is, climate and soil
type determine the maximum potential biomass for a given site, while perturbations
will determine the actual existing biomass (Fig. 1.4). This effect seems to be particu-
larly important in tropical savannas. Whereas areas with high rainfall become for-
ests (too wet to burn) and areas with low rainfall become grasslands, intermediate
woody cover occurs elsewhere, and the degree of canopy cover is maintained by the
intensity, frequency, and severity of the fires present (Staver et al. 2011). It has been
estimated that the percentage of closed fires would approximately double in the
absence of fire (an increase from 27% to 56% of the land surface) (Bond et al. 2005).
Since fires consume biomass, it would appear fair to ask why is the world still
green? Or, in other words, why have fires not entirely deforested the planet?
(Wilkinson and Sherratt 2016). Spatial heterogeneity in fuel and landscapes means
8 1 Introduction

Climatic potential

Edaphic constraints
Biomass

Fire

Actual biomass

Precipitation

Fig. 1.4 Fire interacts with climate and soil to determine actual biomass. In some ecosystems,
herbivores may exert an additional control

that not all areas or ecosystems are equally flammable. Some areas do not support
enough biomass to carry a fire, and others are too wet to burn for a large part of the
year. Oxygen concentration in the atmosphere plays an additional constraint because
if oxygen levels would arose above 25%, flammability of all plant material would
be extremely high and not limited by moisture (Belcher et al. 2010) (see Chap. 3).
Fires often act on the type of vegetation present at the site, but only seldom is veg-
etation completely absent from a site due to fire (Karavani et al. 2018).
Fires affect plants and they may be a selective agent for the physiological traits
they show, as different traits are beneficial under different fire regimes (see Chap.
4). The most basic classification of plants based on their relationship with fire
depends on their regeneration mode (Pausas et al. 2004). Hence, we can observe
plants that resprout after a fire (obligate resprouters), that regenerate from seed
(seeder), that show both traits (facultative resprouters), or that disappear after fire
(fire avoiders). The evolution of many additional traits including bark thickness,
germination cues, serotiny, or canopy architecture, to name a few, has been docu-
mented to be, at least partly, shaped by fire (Tapias et al. 2004; Karavani et al. 2018;
Schwilk 2003).
Wildfires shape soil properties in several different important ways. Fires tempo-
rarily increase nutrient availability through mineralization, but important nutrient
losses may occur through volatilization (Dannenmann et al. 2018). Additionally,
post-fire soil erosion also leads to fertility losses. Soil fertility is also affected by
reductions in soil organic matter and increases in ash content. Mycorrhizae are often
affected negatively by fire, which further hinders the capacity of plants to absorb
nutrients.
Wildfires play a major role in the global C budget, and they are responsible for
the emission of 1–2 Pg C year−1 (van der Werf et al. 2017; Le Quéré et al. 2016),
which represents ~ 10–20% of all emissions (Le Quéré et al. 2016). Additionally,
1.5 Fire in Societies 9

wildfires may serve to maintain biodiversity by increasing landscape heterogeneity


and, consequently, creating a mosaic of ecosystem structures and habitats that serve
as home to a wide variety of organisms when appropriate fire regimes are applied
(Kelly and Brotons 2017). However, fires may also exert negative effects for biodi-
versity, specially under highly anthropogenic fire regimes that lead to mega-fires
(Bowman and Murphy 2010). The effects of fire on the Earth system are discussed
in Chap. 3 and biodiversity consequences in Chap. 10.

1.5 Fire in Societies

Developed societies are scared of fire, and this has the side effect of increasing the
occurrence of catastrophic fires (Fig. 1.5). We conceive the blaze as a destructive
force and, consequently, we seek to put it out as early as it is possible. An unin-
tended consequence of such fire suppression policy is the so-called fire paradox (or
fire-fighting trap), where putting out all fires leads to high accumulations of fuel
loads (Minnich 1983). This is because the fuels that would naturally be consumed
in low-intensity fires accumulate over time as a result of the lack of fire. Such high
fuel loads in turn lead to high-intensity fires that may become beyond extinction
capacity and consume large portions of the land. Thus, while most fires are easily
and promptly extinguished, a few breakaway fires are responsible for most of the
total burned area. The consequence of fearing fires, which is putting them out imme-
diately is leading to larger and more catastrophic wildfires.
Other factors that contribute to large wildfires include the expansion of misman-
aged plantations (in the Mediterranean Basin and Chile), rural exodus and land
abandonment (in the Mediterranean Basin), and the spread of invasive species (in N
America, Chile, and South Africa), along with climate changes which further
increase drought in most fire-prone areas globally. All these factors, linked with fire
suppression, lead to large spatial fuel continuities that, when exposed to dry condi-
tions, are the recipe for large fires. The problem of large wildfires may turn cata-
strophic when it merges with poor urbanization patterns. Areas at the interface
between the wildland and urban areas (WUI), where people live in low-density
settlements and do not manage the wild areas nearby, are creating major problems
for civil protection, particularly in Australia and North America. In contrast, cata-
strophic fires are rare in farming areas with low population densities at the WUI
because of land management and fuel reduction treatments.
Wildfire management has also undergone a transformation in the last decades in
Europe. Starting in the 1960s, linear firebreaks were created with the hope of inter-
rupting fuel continuity. However, large intensity fires are able to “jump” over these
breaks, and consequently, larger areas are now starting to be treated. The question
though is how much of the land needs to be treated to be effective. Wildfires are a
rare occurrence in many areas globally (Table 1.1). Consequently, it would be very
fortuitous that a wildfire encounters one of the treated areas unless, of course, a
large portion of the land is treated. Some recent studies estimate that 15% or more
10 1 Introduction

A natural Human
Action Effect Consequence
phenomenon reaction

Catastrophic
Fear Suppression Fuel loads
fire
Wildfire
Healthy fuel Sustainable
Science Fire mgmnt
loads fire regime

Fig. 1.5 Wildfires are a natural component in many ecosystems worldwide, and we can only learn
to coexist with them. Our reaction to wildfires may either exacerbate or mitigate the fire problem.
When we respond to fire with our emotions, fear takes over and the action will be to put out fires
as quickly as possible. Over time this increases fuel loads, which then pave the way for cata-
strophic fires, further reinforcing the fear we feel over wildfires. We need to break away from this
fear trap. A second reaction to wildfires is to use science to manage wildfires such that we achieve
healthy fuel loads, and in turn, this then leads to a sustainable fire regime

of the land would need to be treated to ensure no large fires occur (Alcasena et al.
2017). The cost of treating such a large area would be unaffordable, and others have
proposed that managing as little as 1% of the land may be enough to avoid large
catastrophic fires (Piñol et al. 2007) for as long as fire prevention measures are
developed in critical and strategic points. While feasible, this beautiful hypothesis
still awaits testing (Chap. 9).

1.6 Scope and Structure of This Book

This book aims at exploring the reciprocal links that exist between physiological
ecology, fuel dynamics, and wildfires. In order to do so, the book has been divided
in three sections. The first section provides an introduction to wildfires. Here we
will discuss the concept of the fire regime, which is of central relevance for under-
standing fire effects on ecosystems (Chap. 2), the effects of fire as an element of the
Earth system (Chap. 3), and how have wildfires shaped the evolution of physiologi-
cal trait (Chap. 4).
The second section develops the concept of pyrophysiology and how it can help
in understanding fire risk, fuel dynamics and development, plant survival after fire,
and vegetation dynamics. Pyrophysiology can help in understanding fire danger by
providing a mechanistic understanding to changes in fuel moisture, as we will
explore in Chap. 5. Fuel development across years and its seasonal dynamics are
discussed in Chap. 6, where we will present the reader with the relevant information
to understand the physiological drivers behind fuel attributes affecting wildfire
spread. The mechanisms explaining how plants succumb after a fire are presented in
Chap. 7. And we will discuss an ecophysiological approach to post-fire vegetation
dynamics in Chap. 8.
References 11

The third section presents the relevance of pyrophysiology from an applied per-
spective. In Chap. 9 we will discuss how can pyrophysiology inform land manage-
ment. Wildfires have become a major socioeconomic problem, and we will discuss
ecophysiological approaches to land management that may mitigate this problem.
We will also try to solve long-standing issues on the relative importance of plant
physiological traits as drivers of landscape flammability and fire behavior. In the last
chapter, Chap. 10, we present the ecophysiological mechanisms required to under-
stand the evolution of wildland fuels under global change.
This book is focused primarily on Mediterranean and European ecosystems and,
to a minor degree, on Australian temperate ecosystems. The physiological mecha-
nisms underlying plant structure and function are rarely biome-specific, and these
contents may thus be of relevance for a broader audience. However, there are many
differences to fire management globally, as well as some fire adaptations that only
appear outside Mediterranean environments, and those may not be discussed here.
While our main focus is the Mediterranean basin, in many parts of the book, we will
draw examples from other parts of the world. Furthermore, Chaps. 2 on fire regimes,
3 on fire and the Earth system, and 10 on wildfires and global change are global in
scope. It is thus expected that this book will be of interest to readers interested in
fire-prone environments globally.

References

Alcasena FJ, Ager AA, Salis M, Day MA, Vega-Garcia C (2017) Optimizing prescribed fire allo-
cation for managing fire risk in Central Catalonia. Sci Total Environ 621:872–885. https://doi.
org/10.1016/j.scitotenv.2017.11.297
Andela N, Morton DC, Giglio L, Chen Y, van der Werf GR, Kasibhatla PS, DeFries RS, Collatz
GJ, Hantson S, Kloster S, Bachelet D, Forrest M, Lasslop G, Li F, Mangeon S, Melton
JR, Yue C, Randerson JT (2017) A human-driven decline in global burned area. Science
356(6345):1356–1362. https://doi.org/10.1126/science.aal4108
Bar A, Michaletz ST, Mayr S (2019) Fire effects on tree physiology. New Phytol. https://doi.
org/10.1111/nph.15871
Belcher CM, Yearsley JM, Hadden RM, McElwain JC, Rein G (2010) Baseline intrinsic flamma-
bility of Earth’s ecosystems estimated from paleoatmospheric oxygen over the past 350 million
years. Proc Natl Acad Sci 107(52):22448–22453. https://doi.org/10.1073/pnas.1011974107
Boer MM, Nolan RH, Resco De Dios V, Clarke H, Price OF, Bradstock RA (2017) Changing
weather extremes call for early warning of potential for catastrophic fire. Earth’s Future
5:1196–1202. https://doi.org/10.1002/2017EF000657
Boer MM, Resco De Dios V, Stefaniak EZ, Bradstock RA (2019) A hydroclimatic model for the dis-
tribution of fire on Earth. Biogeosci Discuss 2019:1–21. https://doi.org/10.5194/bg-2019-441
Bond WJ, Keeley JE (2005) Fire as a global “herbivore”: the ecology and evolution of flammable
ecosystems. Trends Ecol Evol 20(7):387–394. https://doi.org/10.1016/j.tree.2005.04.025
Bond WJ, Woodward FI, Midgley GF (2005) The global distribution of ecosystems in a world
without fire. New Phytol 165(2):525–538
Bond-Lamberty B, Peckham SD, Ahl DE, Gower ST (2007) Fire as the dominant driver of cen-
tral Canadian boreal forest carbon balance. Nature 450(7166):89. https://doi.org/10.1038/
nature06272
12 1 Introduction

Bowman DMJS, Murphy BP (2010) Fire and biodiversity. In: Sodhi NS, Ehrlich PR (eds)
Conservation biology for all. Oxford University Press, Oxford, pp 163–180
Bradstock RA (2010) A biogeographic model of fire regimes in Australia: current and future impli-
cations. Glob Ecol Biogeogr 19:145–158. https://doi.org/10.1111/j.1466-8238.2009.00512.x
Chuvieco E, Yue C, Heil A, Mouillot F, Alonso-Canas I, Padilla M, Pereira JM, Oom D, Tansey
K (2016) A new global burned area product for climate assessment of fire impacts. Glob Ecol
Biogeogr 25(5):619–629. https://doi.org/10.1111/geb.12440
Costafreda-Aumedes S, Comas C, Vega-Garcia C (2017) Human-caused fire occurrence model-
ling in perspective: a review. Int J Wildland Fire 26(12):983. https://doi.org/10.1071/wf17026
Dannenmann M, Díaz-Pinés E, Kitzler B, Karhu K, Tejedor J, Ambus P, Parra A, Sánchez L, Resco
V, Ramírez D, Povoas-Guimaraes L, Zechmeister-Boltenstern S, Kraus D, Castaldi S, Vallejo
A, Rubio A, Moreno J, Butterbach-Bahl K (2018) Post-fire nitrogen balance of Mediterranean
Shrublands: direct combustion losses versus gaseous and leaching losses from the post-fire soil
mineral nitrogen flush. Glob Chan Biol 24:4505–4520
Dickinson MB, Johnson EA (2001) Fire effects on trees. Forest fires: behavior and ecological
effects. Academic Press, San Diego, CA. https://doi.org/10.1016/B978-012386660-8/50016-7
Giglio L, Randerson JT, van der Werf GR (2013) Analysis of daily, monthly, and annual burned
area using the fourth-generation global fire emissions database (GFED4). J Geophys Res
Biogeo 118:317–328. https://doi.org/10.1002/jgrg.20042
Glasspool IJ, Edwards D, Axe L (2006) Charcoal in the early Devonian: a wildfire-derived
Konservat–Lagerstätte. Rev Palaeobot Palynol 142(3):131–136. https://doi.org/10.1016/j.
revpalbo.2006.03.021
Hood SM, Varner JM, van Mantgem P, Cansler CA (2018) Fire and tree death: understanding and
improving modeling of fire-induced tree mortality. Environ Res Lett 13(11):113004. https://
doi.org/10.1088/1748-9326/aae934
Jolly W, Johnson D (2018) Pyro-ecophysiology: shifting the paradigm of live Wildland fuel
research. Fire 1(1):8. https://doi.org/10.3390/fire1010008
Karavani A, Boer MM, Baudena M, Colinas C, Díaz-Sierra R, Pemán J, de Luís M, Enríquez-­
de-­
Salamanca Á, Resco de Dios V (2018) Fire-induced deforestation in drought-prone
Mediterranean forests: drivers and unknowns from leaves to communities. Ecol Monogr
88:141–169
Kavanagh KL, Dickinson MB, Bova AS (2010) A way forward for fire-caused tree mortality pre-
diction: Modeling a physiological consequence of fire. Fire Ecol 6:80–94
Keane RE (2015) Wildland fuel fundamentals and applications. Springer, Cham
Keeley JE, Bond WJ, Bradstock RA, Pausas JG, Rundel PW (2012) Fire in Mediterranean ecosys-
tems- ecology, evolution and management. Cambridge University Press, Cambridge
Kelly LT, Brotons L (2017) Using fire to promote biodiversity. Science 355(6331):1264–1265.
https://doi.org/10.1126/science.aam7672
Knapp AK (1985) Effect of fire and drought on the ecophysiology of Andropogon gerar-
dii and Panicum virgatum in a tallgrass prairie. Ecology 66(4):1309–1320. https://doi.
org/10.2307/1939184
Le Quéré C, Andrew RM, Canadell JG, Sitch S, Korsbakken JI, Peters GP, Manning AC, Boden
TA, Tans PP, Houghton RA, Keeling RF, Alin S, Andrews OD, Anthoni P, Barbero L, Bopp L,
Chevallier F, Chini LP, Ciais P, Currie K, Delire C, Doney SC, Friedlingstein P, Gkritzalis T,
Harris I, Hauck J, Haverd V, Hoppema M, Klein Goldewijk K, Jain AK, Kato E, Körtzinger A,
Landschützer P, Lefèvre N, Lenton A, Lienert S, Lombardozzi D, Melton JR, Metzl N, Millero
F, Monteiro PMS, Munro DR, Nabel JEMS, Nakaoka S-i, O’Brien K, Olsen A, Omar AM,
Ono T, Pierrot D, Poulter B, Rödenbeck C, Salisbury J, Schuster U, Schwinger J, Séférian R,
Skjelvan I, Stocker BD, Sutton AJ, Takahashi T, Tian H, Tilbrook B, van der Laan-Luijkx IT,
van der Werf GR, Viovy N, Walker AP, Wiltshire AJ, Zaehle S (2016) Global Carbon Budget
2016. Earth System Science Data 8(2):605–649. https://doi.org/10.5194/essd-8-605-2016
Marlon JR, Bartlein PJ, Gavin DG, Long CJ, Anderson RS, Briles CE, Brown KJ, Colombaroli
D, Hallett DJ, Power MJ, Scharf EA, Walsh MK (2012) Long-term perspective on wildfires
References 13

in the western USA. Proc Natl Acad Sci USA 109(9):E535–E543. https://doi.org/10.1073/
pnas.1112839109
Minnich R (1983) Fire mosaics in Southern California and Northern Baja California. Science
219:1287–1294
Nelson RF (2001) Water relations of forest fuels. In: Johnson EA, Miyanishi K (eds) Forest fires:
behavior and ecological effects. Academic Press, New York, pp 79–149
Niccoli F, Esposito A, Altieri S, Battipaglia G (2019) Fire severity influences ecophysiological
responses of Pinus pinaster ait. Front Plant Sci 10:539. https://doi.org/10.3389/fpls.2019.00539
Nolan RH, Boer MM, Resco de Dios V, Caccamo G, Bradstock RA (2016) Large scale, dynamic
transformations in fuel moisture drive wildfire activity across South-Eastern Australia. Geophys
Res Lett 43:4229–4238
Nolan RH, Hedo J, Arteaga C, Sugai T, Resco de Dios V (2018) Physiological drought responses
improve predictions of live fuel moisture dynamics in a Mediterranean forest. Agric For
Meteorol 263:417–427. https://doi.org/10.1016/j.agrformet.2018.09.011
Parra A, Moreno JM (2017) Post-fire environments are favourable for plant functioning of
seeder and resprouter Mediterranean shrubs, even under drought. The New Phytologist
214(3):1118–1131. https://doi.org/10.1111/nph.14454
Pate JS, Froend RH, Bowen BJ, Hansen A, Kuo J (1990) Seedling growth and storage characteris-
tics of seeder and resprouter species of Mediterranean-type ecosystems of S.W. Australia. Ann
Bot 65:585–601
Pausas J, Bradstock RA (2007) Plant persistence fire traits along a productivity and disturbance gra-
dient in Mediterranean shrublands of southeastern Australia. Glob Ecol Biogeogr 16:330–340
Pausas JG, Ribeiro E (2013) The global fire–productivity relationship. Glob Ecol Biogeogr
22:728–736
Pausas JG, Bradstock RA, Keith DA, Keeley JE, G.C.T.E. Fire Network (2004) Plant functional
traits in relation to fire in crown-fire ecosystems. Ecology 85:1085–1100
Piñol J, Castellnou M, Beven KJ (2007) Conditioning uncertainty in ecological models: assessing
the impact of fire management strategies. Ecol Model 207:34–44. https://doi.org/10.1016/j.
ecolmodel.2007.03.020
Resco de Dios V, Fellows AW, Nolan RH, Boer MM, Bradstock RA, Domingo F, Goulden ML
(2015) A semi-mechanistic model for predicting the moisture content of fine litter. Agric For
Meteorol 203:64–73
Schwilk DW (2003) Flammability is a niche construction trait: canopy architecture affects fire
intensity. Am Nat 162:725–733
Staver AC, Archibald S, Levin SA (2011) The global extent and determinants of savanna and
forest as alternative biome states. Science 334(6053):230–232. https://doi.org/10.1126/
science.1210465
Tapias R, Climent J, Pardos JA, Gil L (2004) Life histories of Mediterranean pines. Plant Ecol
171(1–2):53–68
Thanos CA, Georghiou K (1988) Ecophysiology of fire-stimulated seed-germination in Cistus
incanus ssp creticus (L) Heywood and Cistus salvifolius L. Plant Cell Environ 11(9):841–849.
https://doi.org/10.1111/j.1365-3040.1988.tb01910.x
van der Werf GR, Randerson JT, Giglio L, van Leeuwen TT, Chen Y, Rogers BM, Mu M, van
Marle MJE, Morton DC, Collatz GJ, Yokelson RJ, Kasibhatla PS (2017) Global fire emis-
sions estimates during 1997–2016. Earth Syst Sci Data 9(2):697–720. https://doi.org/10.5194/
essd-9-697-2017
West AG, Nel JA, Bond WJ, Midgley JJ (2016) Experimental evidence for heat plume-induced
cavitation and xylem deformation as a mechanism of rapid post-fire tree mortality. New Phytol
211(3):828–838. https://doi.org/10.1111/nph.13979
Wilkinson DM, Sherratt TN (2016) Why is the world green? The interactions of top–down and
bottom–up processes in terrestrial vegetation ecology. Plant Ecol Divers 9(2):127–140. https://
doi.org/10.1080/17550874.2016.1178353
Chapter 2
Fire Regimes Across Space

Abstract The effects of fire on landscapes depend on the interaction between the
five components that integrate the fire regime. That is, plants are not adapted to fire
per se, but to a certain fire regime. Similarly, wildfire management depends on the
interaction between the biotic and abiotic features on the site and the fire regime.
The fire regime, in turn, is defined on the basis of the type of fire that occurs in an
area, its intensity and severity, frequency, seasonality, and the mosaic of burned
patches. At global scales, fire regime syndromes may be grouped into five different
pyromes. In this chapter, we explore the notion of the fire regime and its implica-
tions from local to global scales.

2.1 The Fire Regime

The fire regime is a central concept in fire science, and it includes the temporal
and spatial patterns of fire activity. The effects of fires on ecosystems, and its
implications for land management, should be viewed through the lens of the fire
regime. That is, fire frequencies, intensities, size distributions, and seasons follow
certain patterns that vary markedly across biomes, regions, and ecosystem types
(Archibald et al. 2013). Consequently, statements like “this species shows adapta-
tions to fire” are incomplete because different plant species show adaptations to
different fire regimes. Consequently, plant adaptations to a given fire regime may
be maladaptive for a different fire regime. To put an example, the development of
aerial seed banks, which is adaptive in environments where crown fires occur at
decadal intervals, would be maladaptive in areas where fires burn annually
because individuals would die after the second burn, before they had had time to
make seeds.
The fire regime concept has a long tradition in the fire literature (Krebs et al.
2010), and clear definitions started to arise in the 1970s (Gill 1975), which included
the patterns of fire frequency, intensity, and seasonality. In recent years, the fire
regime concept has been expanded to include five components (Keeley et al. 2012),
namely, (1) fire type by fuel layer carrying the fire, (2) fire intensity and severity,
(3) temporal patterns, (4) burn patch size and distribution, and (5) fire seasonality.

© Springer Nature Switzerland AG 2020 15


V. Resco de Dios, Plant-Fire Interactions, Managing Forest Ecosystems 36,
https://doi.org/10.1007/978-3-030-41192-3_2
16 2 Fire Regimes Across Space

We will now discuss each of these components of the fire cycle, and we will end the
chapter by examining global variation in fire regimes. It is important to note that all
these factors covary with each other (Archibald et al. 2013; Murphy et al. 2013).
Thus, they should be considered jointly as the effect of each single factor depends
on interactions with the other factors.

2.2 Fire Type by Fuel Layer Carrying the Fire

The fuelbed complex is composed of three different fuel layers: crown, surface
(understory), and ground fuels. Based upon the fuel layer that carries the fire, we
may thus distinguish between crown, surface, and ground fires. The distinction
between fire types has sometimes been proposed to depend on the fuel layer that is
consumed during the fire (Keeley et al. 2012). However, this definition is problem-
atic because fires that propagate through a layer may exert a substantial impact upon
other layers. That is, a surface fire may consume crown fuel if it has enough inten-
sity. It is thus more clarifying to divide fire types by the layer that carries the fire
along with flame height.

2.2.1 Surface Fires

Surface fires are those where the fire spreads by the vegetation and dead litter in the
understory (up to 2 m above the duff layer). Surface fires have traditionally been
considered as cool, or low-intensity fires, with temperatures only rarely exceeding
400 °C (but note that fire intensity is measured in power, not temperature, as we will
later discuss). However, there is some evidence indicating that the intensity as well
as the severity of surface fires has increased in recent years. As an example, on June
17, 2017, in Pedrógão Grande, Portugal, a surface fire that burned with enough
energy to advance 5000 ha in a single hour (Comissão Técnica Independiente 2017)
led to a complete canopy defoliation across large areas. Fire consumption in surface
fires is thus not restricted to the understory. The front radiative power along with the
convective heat column may provide enough energy to partially or completely defo-
liate the canopy and can be responsible for significant fatalities.
Surface fires occur in grasslands, where there is no forest canopy, or in forests
with a strong vertical separation between the understory and the crown. There are
two major fuel types within the understory, which significantly impact the fire
regime. Understory fuels may be grassy, thus increasing fuel load immediately after
wet periods, or litter fuels, which typically increase in abundance during dry peri-
ods. One of the most common objectives of wildland management in fire-prone
areas is to break the vertical fuel ladder, such that surface fires do not become crown
fires. This is often achieved by thinning small individuals, by eliminating the
2.2 Fire Type by Fuel Layer Carrying the Fire 17

shrubby vegetation such that only herbaceous or annual vegetation may grow, or
through pruning.
Surface fires commonly occur in mountainous regions within Europe or N
America, an area where it is common to encounter species with thick barks that
avoid thermal cambium damage. The natural fire regime in such areas is thought to
naturally prevent the occurrence of crown fires by periodic burning of the under-
story, which naturally prevents the formation of a fuel ladder. However, land aban-
donment and fire suppression, two processes that favor thickening of undergrowth
and therefore vertical fuel continuity, are thought to be largely responsible for trans-
forming the prevailing fire type in many mountainous areas from a surface to
crown fires.

2.2.2 Crown Fires

Crown fires spread through crown fuels, which are those occurring 2 m above the
duff layer. A common distinction between surface and crown fires is thus whether
or not the flame is higher than 2 m. There are three different types of crown fires,
and they show markedly different characteristics as well as ecosystem effects (Van
Wagner 1977). Passive or dependent crown fires refer to the torching of individual
trees or of a group of trees. Tree torching occurs when the intensity of a surface fire
is high enough to ignite the canopy, but there is not enough fuel build-up in the
canopy to allow its propagation. When canopy fuel loads are not enough for fire
propagation, these crown fires thus depend on surface fire conditions. Tree torching
or passive crown fires may occur when there are local accumulations of ladder fuels
and are more common in areas with a low density of trees (i.e., open forests or
savannas) or in areas that have been defoliated recently and lack horizontal fuel
continuity. Among the different types of crown fires, these often show the lowest
intensities.
Active crown fires occur when the fire spreads across a combination of crown and
surface fuels. This is the case when the intensity of the surface fire is high enough
to ignite the canopy and, at the same time, there is enough fuel build-up in the can-
opy to allow its spread through the stand. Active crown fires often exceed extinction
capacity, and these fires can thus not be suppressed by current firefighting
capabilities.
Independent crown fires burn at higher intensities than the previous two fire
types. They occur when the intensity of the surface fire is high enough to reach the
canopy, there is enough canopy fuel built-up to allow its propagation, and the energy
released by the crown fire is high enough to maintain it independently from the
surface fire. Independent crown fires may occur under very high fuel build-ups in
shrublands or forests, and they often also require the presence of very strong winds
or a topography that favors such extreme fire behavior.
18 2 Fire Regimes Across Space

Crown fires are typical of Mediterranean environments, and some adaptations to


this fire type include the development of seed banks, either in the canopy or in the
soil, or resprouting capabilities. It has been argued that some of these species that
have seed banks and are therefore adapted to crown fires may actually favor the
development of crown fires, either by creating ladder fuels or by emitting highly
flammable volatile substances. However, while appealing, demonstrating the valid-
ity hypothesis has proven challenging (see Chap. 4).

2.2.3 Ground Fires

Ground fires are typically of low intensity but high severity as they spread by smol-
dering combustion of the duff, roots, and soil organic matter. These fires tend to
burn very slowly due to the low oxygen supply, and consequently, the combustion
rate of the organic matter, roots, and other fuels is very high. Ground fires are com-
mon in poorly aerated soils with very high accumulations of organic matter such as
in peats or mires. These fires are very difficult to extinguish. Their main relevance
to aboveground fire is that they may, sometimes, serve as ignition point for a surface
or crown fire, for instance, if the fire is not completely put out during extinguishing
and mop-up operations. Ground fires are particularly problematic in areas where
peatlands have been drained and the land use has changed to plantations, as will be
discussed in Chaps. 3 and 10.

2.3 Intensity and Severity

Fire intensity and severity are terms that have been used interchangeably in the past,
but they do have very different meanings (Keeley 2009). Fire intensity refers to the
amount of energy released during the fire, while severity indicates the ecological
impact of the fire in terms, generally, of biomass loss. Just like studies on drought
need to provide an indicator of the intensity of the water stress, studies on fire also
need to provide an estimate of its intensity. It is very difficult to understand fire
effects unless the intensity, either measured directly or through some proxy, is given.

2.3.1 Fire Intensity

The definitions of fire intensity have been changing over time as a result of improved
physical representations of fire behavior as well as of technological advancements.
The earliest definition of fire intensity (I) was provided by Byram (Fig. 2.1):

I = HBR (2.1)
2.3 Intensity and Severity 19

Fig. 2.1 Definitions of fire intensity. (a) Fireline intensity is the energy released by a 1-m-wide
strip of the flame front. (b) Reaction intensity is the energy released in 1 m2 of the flame front.
Photo by José Carlos López Sánchez

where H is the heat yield of the fuel (KJ/Kg), B the fuel consumed during the fire
(Kg/m2), and R the rate of spread (m/s). Byram’s definition was defined to quantify
the intensity of the fire front, and it is thus expressed as kW/m. As a result, Byram’s
intensity is often referred to as fireline intensity. Note that W is often used instead of
B to refer to fuel consumed, but here we use B to avoid confusion with watt.
Following physical conventions, intensity should be defined as the energy per
unit area and per unit time, which is measured as W/m2. Using Rothermel’s reaction
intensity, fire intensity was redefined as:

I = I r tr R (2.2)

where Ir indicates reaction intensity and tr the residence time.


The development of remote sensing and geographical information systems has
allowed for satellite estimates of fire intensity, which often use the radiant energy
(fire radiative power, FRP, in W per unit area) using thermal imagery in the infrared
spectrum (Johnston et al. 2017).
The use of each definition of intensity will vary with the objectives. Equation
(2.2), for instance, is largely limited to modeling efforts. From an operational per-
spective, and for on-ground studies on fire intensity and its effects on the environ-
ment or infrastructure, Byram’s fireline intensity is most commonly used. It indicates
the energy released by a 1-m-wide fire band across the fireline, from the rear to the
leading edges (Fig. 2.1). I varies mostly with B and R, since H is largely constant
across fuel types (Van Wagner 1972). The interplay between B and R in driving I
may be seen in Fig. 2.2. Grassy fuels tend to exhibit high R but low W and, conse-
quently, low I. Heavy fuels such as slash residues often show high to mid R but high
B, consequently showing larger I than cured grasslands. Shrubland and closed for-
ests will show the highest R and B and, consequently, the highest I (Alexander and
20 2 Fire Regimes Across Space

F
in ireli
te ne

1.5
ns
ity

Rate of spread (m s−1)

70
1.0

00
0
11
0.5

07
5
52

17
65

02
0.0

0 1 2 3 4 5 6
Fuel consumed (kg m −2
)
Fig. 2.2 Values of fireline intensity (kW/m), as a function of rate of spread and of fuels consumed,
that are critical for operational purposes across shrublands (dashed line, Fernandes et al. 2000) and
pine litter with grasses (straight line, Byram 1959). Yellow, purple, and red indicate the intensities
that lead to flames of 1.5, 2.5, and 3.5 m, respectively. Numbers indicate fireline intensity values
for shrublands. A higher fireline intensity of 70,000 kW/m, where firestorms may occur, is also
included

Cruz 2019). Landscape-scale patterns of variation in fire intensity thus result from
spatial variation in fuel loads.
In practice, I is often inferred from the length of the flame (lf) and from the height
of the stem char (hc) or of the crown scorch (hs). Byram (1959) already proposed an
empirical equation that relates I to lf, but such relationships are not universal as they
vary across fuel types and fire spread patterns. hc and hs are useful post-fire indices
of I. The rationale behind such relationships is that, on the one hand, hc directly
reflects lf (an assumption better suited for nonfibrous bark species, as char height
may vary after fire in flammable barks), and on the other hand, hs is also the result
from the interplay between fire radiation and the lethal tissue temperature. The use
of hc and hs as indicators of I assumes that the flame is perfectly perpendicular to the
ground, and consequently, steep slopes, windy conditions, or burning conditions
increase the uncertainty on I estimations from hc and hs. In shrublands, additionally,
the diameter of finest twig in burned individuals also correlates fire intensity, but
this method is site- and species-specific (Moreno and Oechel 1989).
Fireline intensity estimates provide information that will be relevant for firefight-
ing and, in consequence, for wildland planning. Direct fire attacks with manual
tools are often possible under lf below 1.5 m, and the upper limit for direct attacks
is considered to be at 2.5 m (Molina et al. 2009). When fires present lf between 2.5
and 3.5 m, serious problems for direct attacks occur, and crowning and spotting
become increasingly likely. This information can then be used for wildfire
2.3 Intensity and Severity 21

management. That is, fuel reduction treatments are often performed to reduce the
potential intensity of a fire. Using available information on the relationship between
lf and I that is available for different vegetation types, we can then decide the inten-
sity of the fuel reduction treatment that will be necessary to achieve the target goals.
Figure 2.2 shows the different combinations of R and W that lead to such critical
values of lf and I for direct attack and crowning for two different vegetation types:
pine litter with grass understory (Byram 1959) and shrublands (Fernandes et al.
2000). This type of models may then be used to design fuel reduction treatments
such that I does not reach critical limits. A particularly important limit to avoid is
I > 70,000 kW/m, as firestorms then become possible (Alexander and Cruz 2019).

2.3.2 Fire Severity

Fire severity indicates the impact that the fire has had upon a feature of interest.
Definitions of fire severity may thus differ between our objectives. However, bio-
mass loss has been proposed as a relatively simple and quantifiable metric of fire
severity (Keeley 2009). It is important to note that fire severity is not always related
to fire intensity (Fig. 2.3). Extreme examples of this are ground fires, where mortal-
ity is often very high although the fire is flameless.
One of the most common indicators of fire severity is tree mortality or losses in
photosynthetic tissues, which can be inferred from wildland inventories and also
from remote sensing techniques. Fire severity through remote sensing is often
inferred by comparing the values of some index before and after the fire. In principle,
preferred indices should be those that are directly related to biomass loss, such as leaf

Fig. 2.3 Difference between intensity (I) and severity (S). The area under high intensity and high
severity (HI HS) was a crown fire (indicated by leaves burned) and all trees died. The area under
low intensity and high severity (LIHS) was a low-intensity surface fire where all individuals died
from radiation (all leaves brown). The area under low intensity and low severity (LI LS) was a
surface fire with little impacts on the canopy (leaves green). Note how burn intensity and severity
changed in the landscape while community composition did not, indicating that landscape flam-
mability does not solely depend on vegetation traits
22 2 Fire Regimes Across Space

area index (LAI), or the change in the fraction of absorbed photosynthetically active
radiation (faPAR) (Boer et al. 2008). However, the difference between pre- and post-
fire normalized burn ratio (ΔNBR), which compares values in the near-infrared
(NIR) and middle-wave infrared (MIR) bands, has been most commonly used:

NBR = ( NIR − MIR ) / ( NIR + MIR ) (2.3)

ΔNBR is an empirical indicator of fire severity, and it needs to be calibrated


against on-ground data before its application may be validated (an advice that is
generally applicable to all remote sensing indices). Studies on NBR as an indicator
of fire severity often report mixed results. For example, Murphy et al. (2008) studied
the validity of the ΔNBR to quantify the severity of the boreal forests from Alaska
and observed how the degree of correlation with on-ground forest inventory data
varied markedly across sites (R2 = 0.11–0.64). Part of the problem underlying such
poor correlation is that ground measurements of fire severity are not always designed
to validate the spectral response recorded through a change in NBR. More recently,
different approaches based on machine learning algorithms have been proposed,
and they show promising results (Collins et al. 2018).
There are different predisposing factors that may enhance fire severity. The
degree of environmental stress prior to the fire, for instance, could increase fire
severity. Fire severity also varies markedly across vegetation types, and it also
depends on the time since the last fire. It is important to note that fire severity does
not indicate long-term ecological or social impacts. Fire severity often refers to the
degree of biomass loss and separate studies or metrics need to be developed to study
socio-ecological processes.

2.4 Temporal Patterns

Wildfires occur at a given frequency, which varies from yearly in subtropical grass-
lands and savannas to hundreds of years in alpine environments or in very wet for-
ests (Table 2.1). Such diversity of fire return intervals has led to very marked
regeneration strategies and fire adaptations (see Chap. 4).
There are different forms in which we can quantify the temporal pattern of fires
(Oddi 2018). Fire frequency is the number of fire occurrences at a given site per unit
time (fires year−1); the fire return interval indicates the average number of years
between fires at a site (years); and the fire cycle or rotation period represents the
time required to burn an area of equivalent size to that of interest. The difference
between each of these three aspects is subtle, but each responds to a slightly differ-
ent question. For example, fire frequency responds to “how many fires have occurred
in a location?”, fire return interval to “how often do fires occur in an area?”, and fire
rotation to “how long does it take to burn this area? (Oddi 2018).
2.4 Temporal Patterns 23

Table 2.1 Fire return interval for different world regions


Region Fire return interval (year)
Boreal N America 393
Temperate N America 271
Central America 78
N Hemisphere South America 54
S Hemisphere South America 54
Europe 502
Middle East 861
Northern Hemisphere Africa 8
Southern Hemisphere Africa 5
Boreal Asia 158
Central Asia 79
Southeast Asia 43
Equatorial Asia 103
Australia and New Zealand 14
Data from van der Werf et al. (2017)

The calculation of fire frequency is straightforward from the number of fire


events occurred at a given site divided by the number of years over that period. The
fire return interval is simply the reverse of fire frequency, and both can be assessed
either at a point or at an area. The rotation period (RP) or fire cycle is only available
for an area, and it may be calculated as:

p
RP = (2.4)
BA / TA

where p indicates the period (years), BA is the burned area (ha), and TA is the total
area of interest (ha) (Table 2.1).
Archibald et al. (2013) conducted a global assessment of fire regime attributes,
and they distinguished three types of fire return intervals: frequent (1–4 years),
intermediate (6–19 years), and rare (>50 fires) fire intervals. Interestingly, in their
global analysis, they identified that the intermediate fire return interval was only
present in anthropogenic fire regimes. That is, “natural” fire regimes included either
frequent or rare fire return intervals, whereas intermediate fire return intervals were
only present in areas where humans had intensively modified the fire regime (more
details in Sect. 2.7).
The temporal pattern of wildland fires may be inferred through a variety of meth-
ods. At long temporal scales, but poor resolution, changes in fire frequency may be
assessed by charcoal and pollen deposits (Scott et al. 2014). These estimates are
often biased towards assessments of the fire regime at high-elevation or high-­latitude
lakes. Fire frequencies over the past few centuries may be quantified from fire scars
in dendrochronological analyses. These analyses are restricted to forest areas: (1)
with low severity surface fires or other fire types that do not lead to a stand
24 2 Fire Regimes Across Space

replacement and (2) old growth stands or areas where trees survive for long enough
to develop meaningful chronologies that are not affected by management (Swetnam
1993). Fire frequencies or rotation periods may also be inferred from data on fire
inventories, coupled with examination of old newspapers. Fire inventories are usu-
ally available for a few decades and up to a century, and local newspapers allow to
complement this information, particularly in areas where fire inventories are more
recent (Pausas and Fernández-Muñoz 2011). In some cases, however, information
for a few centuries may be inferred from local archives (Camarero et al. 2019).
Information on fire return intervals is also available through remote sensing, but
satellite data is only available for the last few decades, and these fire frequency
estimates are thus mostly suitable for areas with frequent or intermediate fire
frequencies.

2.5 Seasonality

Fire seasonality is associated with the time when the fuel becomes available. In for-
est or shrubland areas, where fuel build-up is high all year round, large wildfires
occur when the fuel is dry enough to burn (see Chap. 1). This may occur in the sum-
mer in the Mediterranean basin or in the autumn in California, but also in the winter
or spring in areas dominated by tropical or monsoonal climates. The effect of fire
seasonality is not yet properly understood. It has been proposed that prescribed
burnings, for instance, may help in restoring the natural fire rotation period.
However, such prescribed burns are often conducted off-season, during a time of the
year when fire danger is not extreme that allows for burn control. There is some data
that indicates how burning at different times of the year exerts different impacts on
post-fire regeneration (Moreno et al. 2011), but more studies are necessary to fully
understand the effects of off-season fires.
The fire season may, in some instances, follow a bimodal pattern. This is the
case, for instance, of N Spain, where peaks in fire occurrence occur in August, the
driest month of the year, but also in winter. Winter fires in this area are often associ-
ated with traditional management activities, which favor pastures at the expense of
forests, and they may sometimes escape. Bimodal fire seasons may thus result from
anthropogenic activities.
Changes in fire seasonality are one of the most noticeable effects of climate
change on the fire regime (Jolly et al. 2015). One the one hand, the start of the fire
season has been reported to advance in recent years in Mediterranean ecosystems
due to the increase in temperatures which consequently favors fuel drying. On the
other hand, there has been an increase in winter wildfires in many European coun-
tries, which would appear to be linked with changes in North Atlantic Oscillation.
The net result of these changes is that the fire season has been increasing over the
last few years. Some European regions are now experiencing a bimodal pattern of
fire activity, with peaks in the summer and also in the winter, and further tempera-
ture increases could potentially lead to a year-around fire season.
2.6 Mosaic 25

2.6 Mosaic

The last aspect of the fire regime is related to the burn mosaic. Fire size distributions
have changed dramatically as a result of the fire suppression policy, and the vast
majority of wildfires are now very small (<50 ha). However, large wildfires (those
with >1000 ha), while rare, are responsible for burning over disproportionately
large areas (Fig. 2.4). In some Western European countries, large wildfires are
<0.2% of all the wildfires, but they are responsible for burning as much as 50% of
the total burned area. This process is known as the paradox of extinction or the fire-
fighting trap. A fire suppression policy, which is very effective at suppressing 99.8%
of all wildfires, may lead to large accumulation of fuel that, in turn, favors the
appearance of large wildfires that consume a large proportion of the area.
Large wildfires are also favored by the continuity of available land under drought.
As water scarcity increases, the number of wildland patches that are available to
burn increases, and in consequence, there is a higher connectivity of dry patches of
land, favoring the occurrence of large wildfires (Caccamo et al. 2012).
26190
Number of forest fires
75
25
0

100 200 300 400 500 600 700 800 900 1000 50001000015000
Fire size (ha)
250000
Cumulative burned
forest area (ha)
100000
0

0 5000 10000 15000

Forest fire size (ha)

Fig. 2.4 Data on wildfire size and number in NE Spain (Catalonia) according to Spanish National
Fire Statistics 1973–2015. Large wildfires burning over 1000 ha are rare (upper panel), but they are
responsible for a large proportion of total burned area (lower panel)
26 2 Fire Regimes Across Space

2.7 Pyromes

To characterize global variation in fire regimes, Archibald et al. (2013) examined


the patterns of covariation between fire size, frequency, intensity, season, and extent
globally. They observed that these fire regime attributes do not vary randomly over
space and that only certain combinations occur. More specifically, they observed
that global variation in fire regimes could be categorized as belonging to one of five
different combinations of fire syndromes, which they called pyromes (Fig. 2.5,
Table 2.2).
Grasslands and shrublands, ecosystems with the capacity for prolific and rapid
resprouting, show frequent fires, and differences in their fire intensity appeared to
differ between Australian grasslands, where fires are intense and large (the frequent-­
intense-­large, FIL pyrome), and African grasslands, where fires are cool and small
(frequent-small-cool, FCS pyrome).
Conifer forests and Mediterranean vegetation experience fires rarely but that
burn at high intensity and over large patches of land (rare-intense-large, RIL
pyrome). This indicates that these ecosystems experience crown fires, and their
regeneration will thus be linked to serotiny or other seed banks. Fires that are rare-­
cool-­small (RCS pyrome) are interspersed with the RIL pyrome, and they are also
frequent at additional biomes such as rainforests or temperate and montane grass-
lands. Understanding why a given biome shows both RCS and RIL pyromes is not
well understood. Some authors propose that plant traits, such as the degree of self-­
pruning, could explain for instance why some boreal forests burn at high-intensity
crown fires, whereas others only burn at low-intensity surface fires (Rogers
et al. 2015).

50°N
Latitude

50°S

100°W 0° 100°E
Longitude

Fig. 2.5 Global distribution of five pyromes: FIL in yellow, FCS in orange, RIL in green, RCS in
pink, and ICS in blue. Pyromes were originally described by Archibald et al. (2013), but here we
use the representation made in Boer et al. (2019)
References 27

Table 2.2 Fire regime attributes in each of the five pyromes: frequent-intense-large (FIL),
frequent-cool-small (FCS), rare-intense-large (RIL), rare-cool-small (RCS), and intermediate-­
cool-­small (ICS)
FIL FCS RIL RCS ICS
% burned 14 9 1 0 0
area
Fire return 3 1 >50 >50 12
interval
(year)
Maximum 473 197 476 187 224
intensity
(fire
radiative
power, MW)
Max fire size 414 25 83 4 9
(km2)
Length fire 4 3 2 1 3
season (mo)
Biome Tropical Tropical Temp and boreal Rainforests, All
grasslands grasslands conifer forest, temperate mixed and biomes
and and Mediterranean boreal forests,
shrublands shrublands and xeric temperate and
and flooded vegetation montane grasslands,
grasslands Mediterranean and
xeric vegetation
Redrawn from Archibald et al. (2013)

The fifth pyrome is characterized by intermediate-cool-small (ICS pyrome) fires,


and it is widespread across the globe but particularly present in areas where agricul-
ture and deforestation are present. The high prevalence of the ICS pyrome hints
towards an anthropogenic origin. That is, this fire regime likely results from human
action, which leads to small-sized fires with reduced intensity. This is consistent
with the effects of a fire suppression policy stated earlier in the chapter.

References

Alexander ME, Cruz MG (2019) Fireline intensity. In: Manzello SL (ed) Encyclopedia
of Wildfires and Wildland-Urban Interface (WUI) fires. Springer, pp 1–8. https://doi.
org/10.1007/978-3-319-51727-8_52-1
Archibald S, Lehmann CER, Gómez-Dans JL, Bradstock RA (2013) Defining pyromes and global
syndromes of fire regimes. Proc Natl Acad Sci 110(16):6442–6447. https://doi.org/10.1073/
pnas.1211466110
Boer MM, Macfarlane C, Norris J, Sadler RJ, Wallace J, Grierson PF (2008) Mapping burned areas
and burn severity patterns in SW Australian eucalypt forest using remotely-sensed changes
in leaf area index. Remote Sens Environ 112(12):4358–4369. https://doi.org/10.1016/j.
rse.2008.08.005
28 2 Fire Regimes Across Space

Boer MM, Resco De Dios V, Stefaniak EZ, Bradstock RA (2019) A hydroclimatic model for the
distribution of fire on Earth. Biogeosci Discuss. https://doi.org/10.5194/bg-2019-441
Byram GM (1959) Combustion of forest fuels. In: Davis KP (ed) Forest fire: control and use.
McGraw-Hill, New York, pp 61–89
Caccamo G, Chisholm LA, Bradstock RA, Puotinen ML (2012) Using remotely-sensed fuel con-
nectivity patterns as a tool for fire danger monitoring. Geophys Res Lett 39(1):L01302. https://
doi.org/10.1029/2011GL050125
Camarero JJ, Sangüesa-Barreda G, Pérez-Díaz S, Montiel-Molina C, Seijo F, López-Sáez JA
(2019) Abrupt regime shifts in post-fire resilience of Mediterranean mountain pinewoods are
fuelled by land use. Int J Wildland Fire. https://doi.org/10.1071/wf18160
Collins L, Griffioen P, Newell G, Mellor A (2018) The utility of random forests for wildfire sever-
ity mapping. Remote Sens Environ 216:374–384. https://doi.org/10.1016/j.rse.2018.07.005
Comissão Técnica Independiente (2017) Relatorio: Análise e apuramento dos factos relativos
aos incêndios que ocorreram em Pedrogão Grande, Castanheira de Pera, Ansião, Alvaiázere,
Figueiró dos Vinhos, Arganil, Góis, Penela, Pampilhosa da Serra, Oleiros e Sertã, entre 17 e 24
de junho de 2017. Assambleia da República
Fernandes PM, Catchpole WR, Rego FC (2000) Shrubland fire behaviour modelling with micro-
plot data. Can J For Res 30:889–899
Gill AM (1975) Fire and the Australian flora: a review. Aust For 38:4–25
Johnston JM, Wooster MJ, Paugam R, Wang X, Lynham TJ, Johnston LM (2017) Direct estimation
of Byram’s fire intensity from infrared remote sensing imagery. Int J Wildland Fire 26(8):668.
https://doi.org/10.1071/wf16178
Jolly WM, Cochrane MA, Freeborn PH, Holden ZA, Brown TJ, Williamson GJ, Bowman DMJS
(2015) Climate-induced variations in global wildfire danger from 1979 to 2013. Nat Commun
6(1):7537. https://doi.org/10.1038/ncomms8537
Keeley JE (2009) Fire intensity, fire severity and burn severity: a brief review and suggested usage.
Int J Wildland Fire 18:116–126
Keeley JE, Bond WJ, Bradstock RA, Pausas JG, Rundel PW (2012) Fire in Mediterranean ecosys-
tems- ecology, evolution and management. Cambridge University Press, Cambridge
Krebs P, Pezzatti GB, Mazzoleni S, Talbot LM, Conedera M (2010) Fire regime: history and
definition of a key concept in disturbance ecology. Theor Biosci 129(1):53–69. https://doi.
org/10.1007/s12064-010-0082-z
Molina D, Blanco J, Galán M, Pous E, Garcia J, Garcia D (2009) Incendios forestales: fundamen-
tos, lecciones aprendidas y retos del futuro. Ediciones AIFEMA, Granada
Moreno JM, Oechel WC (1989) A simple method for estimating fire intensity after a burn in
California chaparral. Acta Oecol 10(1):57–68
Moreno JM, Zuazua E, Pérez B, Luna B, Velasco A, Resco De Dios V (2011) Rainfall patterns
after fire differentially affect the recruitment of three Mediterranean shrubs. Biogeosciences
8(12):3721–3732
Murphy KA, Reynolds JH, Koltun JM (2008) Evaluating the ability of the differenced Normalized
Burn Ratio (dNBR) to predict ecologically significant burn severity in Alaskan boreal forests.
Int J Wildland Fire 17:490–499
Murphy BP, Bradstock RA, Boer MM, Carter J, Cary GJ, Cochrane MA, Fensham RJ, Russell-­
Smith J, Williamson GJ, Bowman DMJS (2013) Fire regimes of Australia: a pyrogeographic
model system. J Biogeogr 40:1048–1058. https://doi.org/10.1111/jbi.12065
Oddi FJ (2018) Fire Regime. In: Manzello SL (ed) Encyclopedia of wildfires and Wildland-urban
Interface (WUI) fires. Springer, pp 1–12. https://doi.org/10.1007/978-3-319-51727-8_73-1
Pausas JG, Fernández-Muñoz S (2011) Fire regime changes in the Western Mediterranean Basin:
from fuel-limited to drought-driven fire regime. Clim Chang 110(1–2):215–226. https://doi.
org/10.1007/s10584-011-0060-6
Rogers BM, Soja AJ, Goulden ML, Randerson JT (2015) Influence of tree species on continental
differences in boreal fires and climate feedbacks. Nat Geosci 8:228–234
References 29

Scott AC, Bowman DMJS, Bond WJ, Pyne SJ, Alexander ME (2014) Fire on Earth: an introduc-
tion. Wiley-Blackwell, West Sussex
Swetnam TW (1993) Fire history and climate change in giant sequoia groves. Science
262(5135):885–889. https://doi.org/10.1126/science.262.5135.885
Van Wagner CE (1972) Heat of combustion, heat yield and fire behaviour. Canadian Forestry
Service, Petawawa Forest Experiment Station, Information report PS-X- 35, Chalk River
Van Wagner CE (1977) Conditions for the start and spread of a crown fire. Can J For Res 7:23–34
van der Werf GR, Randerson JT, Giglio L, van Leeuwen TT, Chen Y, Rogers BM, Mu M, Van
Marle MJE, Morton DC, Collatz GJ, Yokelson RJ, Kasibhatla PS (2017) Global fire emis-
sions estimates during 1997–2016. Earth Syst Sci Data 9 (2):697–720. https://doi.org/10.5194/
essd-9-697-2017
Chapter 3
Fire as an Earth System Process

Abstract This chapter reviews the role of wildfires as an essential element of the
Earth system. After a brief explanation of systems theory, we will discuss how fire
regulates the oxygen cycle and climate. We will then travel across the Earth’s his-
tory to understand how wildfires have shaped the Earth as we know it today. We will
cover how fire activity has varied across geological scales and how different pro-
cesses, including mass extinctions, have been affected by wildfire activity. We will
then review fire-human interactions. Wildfires have served as a powerful tool for
human alteration of landscape structure. We will provide examples of how humans
have modified African, Australian, American, and European landscapes through the
use of fire. We will finally discuss the cost of fire to human lives, as the number of
fire-induced fatalities from smoke is on the rise. Although wildfires are often per-
ceived as a major environmental problem, their effect over the Earth system makes
them an essential element for life.

3.1 The Earth as a System and the Role of Fire

Wildfires are a biosphere-atmosphere phenomenon (Fig. 3.1), with cascading impli-


cations for other sub-systems such as the hydrosphere or, sometimes, the geosphere.
Over short time scales, wildfires affect the interactions between ecosystem structure
and community trait compositions along with energy and greenhouse gas balance
(Fig. 3.1). Over longer time scales, wildfires affect interactions between the evolu-
tion of physiological traits, leading to speciation events, and the oxygen and P
cycles (Fig. 3.1).
Wildfire effects also include a human dimension. Over short time scales, wild-
fires may lead to mortality from the flames or smoke ingestion due to the heat and
particulate matter emissions. Over intermediate time scales, wildfires have served as
a tool for land management that has led to altering the climatic system. Over longer
time scales, wildfires have also contributed to human evolution (Fig. 3.1).
Here we will first discuss how wildfires have shaped the atmosphere. That is, we
will cover how wildfires regulate oxygen concentrations and the climate and water
cycles. We will then describe the changes in fire activity across geological periods.
Next we will move to the human dimension to explain how humans modified

© Springer Nature Switzerland AG 2020 31


V. Resco de Dios, Plant-Fire Interactions, Managing Forest Ecosystems 36,
https://doi.org/10.1007/978-3-030-41192-3_3
32 3 Fire as an Earth System Process

Fig. 3.1 Fire is a phenomenon at the interface between the biosphere and the atmosphere, and it
exerts important consequences for the Earth system across diverse spatiotemporal scales. These
processes also affect to, and are modified by, our species, which is part of the biosphere. In these
triangles, fire is pictured at the top, the biosphere to the left, and the atmosphere to the right

l­andscapes through the use of fire. In the last section, we will present some of the
challenges that wildfires present for humans in the Anthropocene. The effects of
wildfire on the biosphere, including trait evolution and ecosystem structure, will be
covered between Chaps. 3 and 7.

3.1.1 Fire and Oxygen

The role of fire as an element of the Earth system is not a recent one. In fact, wild-
fires are thought to have played a role in stabilizing atmospheric O2 concentrations
at the current level of 21% (Watson et al. 1978; Berner et al. 2003). For instance,
one of the main mechanisms why O2 concentrations are currently stable is related to
the burial of organic carbon. Higher rates of organic carbon burial diminish the
oxidation or organic matter and thus increase O2 concentrations. The amount of O2
that is released through this pathway is relatively small, but large enough to stabilize
O2 concentrations over geological time scales.
Wildfires affect the burial of organic C in different ways, creating both positive
and negative feedbacks over atmospheric O2 concentrations. A positive feedback
occurs through the production of charcoal. Charcoal is very recalcitrant, and it shows
3.1 The Earth as a System and the Role of Fire 33

a large resistance to decomposition. Increasing charcoal formation, which results


from fire activity, thus increases the burial of organic carbon. Another positive feed-
back on atmospheric oxygen occurs through erosion, a process that is enhanced by
wildfires, as it increases sedimentary deposits of organic carbon (Berner et al. 2003).
However, wildfires also exert negative feedbacks over atmospheric O2 concentra-
tions. On the one hand, wildfires consume biomass which then diminishes the
amount of organic C burial. On the other hand, fire exerts a negative feedback over
atmospheric O2 via perturbations of the P cycle (Kump 1988). That is, wildfires
diminish plant cover, which then decreases the weathering of phosphate minerals.
Such a decrease in phosphate weathering could reduce the input of dissolved P into
the oceans, subsequently lowering the amount of P in the ocean and, consequently,
limiting organic production, which would then reduce organic burial and, conse-
quently, O2 production (Berner et al. 2003).
Oxygen levels also affect the sensitivity of fire spread to fuel moisture (Fig. 3.2).
The moisture of the wildland fuel limits wildfire propagation under current O2 concen-
trations such that fires are rare when fuel moisture content is above 15% (Fig. 3.2).
However, such moisture limitation diminishes as oxygen concentrations raise (Fig. 3.2)
(Watson et al. 1978; Belcher et al. 2010). Some authors have argued that atmospheric
oxygen concentrations may have reached the level of 35% at some moment during our
geological history. However, others consider a more likely upper limit to sit at 25%
because, otherwise, the Earth would be too flammable (Lenton 2016) and there would
be a large increase in fire propagation that would then create a negative feedback on
atmospheric O2 levels. Whereas this debate is not yet resolved, it clearly indicates how
the effects of fire transcend short-term effects on vegetation and properties as they also
affect biogeochemical cycling, climate, and atmospheric chemistry.

Fig. 3.2 Fire spread rate increases with oxygen concentration and decreases with fuel moisture.
Fires cannot occur when oxygen levels are below 15% and fuel moisture limitation declines par-
ticularly at oxygen concentrations higher than 25%. (Reproduced with permission from (Watson
and Lovelock 2013))
34 3 Fire as an Earth System Process

3.1.2 Fire-Climate Feedbacks

Fire activity and climate are interlinked. On the one hand, climate affects fire activ-
ity by altering the structure and availability of fuel (Gil-Romera et al. 2014). On the
other hand, wildfires affect the climatic system. This is because of the effects of fire
on the carbon cycle as well as its effects over the energy balance (Fig. 3.3).
Currently, land ecosystems assimilate 25% of all fossil fuel emissions. However,
the future of such carbon sink is unclear. Some climate models project that the
strength of the land C sink could decrease during the current century (Friedlingstein
et al. 2014). If true, this would lead to an amplification of climate change. This is
because if the land C sink diminishes, then a larger portion of emissions will remain
in the atmosphere, thus contributing to climate change. Furthermore, vast amounts
of C are stored globally in soils: returning such C to the atmosphere as CO2 would
further amplify climate change. Wildfires could thus affect the strength of the land
C sink and atmospheric CO2 concentrations in at least two ways: by reducing CO2
assimilation as well as by reducing the carbon stocks in the soils. Additionally, a
fraction of the organic matter that is combusted during the fire is transformed into
highly recalcitrant pyrogenic carbon. Pyrogenic carbon formation partly counters
fire-induced emissions, because this compound is highly resistant to degradation.
In addition, physiognomic changes brought by wildfires alter the energy balance,
by altering albedo and the ratio between latent and sensible heat. Furthermore,

Pyrogenic Carbon
+

+ C stocks -
C Emissions +
-
- - Warming
- C sinks
Precipitation
Wildfires +
- -
ET
- Water-quality

Run-off -
-
+
+ Erosion

Albedo

Fig. 3.3 Fire affects the climate system by altering the carbon and water cycles as well as albedo
3.1 The Earth as a System and the Role of Fire 35

smoke particles interact with water vapor and reduce precipitation. Wildfires also
change the hydrological cycle and may lead to important soil losses through
increased erosion (Fig. 3.3).

3.1.2.1 Fire Reduces CO2 Assimilation

Fire is currently used as a tool for deforestation. This is particularly the case of the
Amazonian Basin. Tropical forests have evolved in largely fire-free environments,
and their trees thus lack the adaptations to survive the fire (Bierregaard et al. 2001).
When a fire occurs in the Amazon, it burns under low intensity but high severity.
That is, its flames are low, often shorter than the knee, but they advance very slowly,
maybe 100–150 m per day. Their slow advancement leads to a very complete com-
bustion of the tree trunk that girdles and, consequently, kills those trees with thin
barks or low diameters (Cochrane and Schulze 1998). After a fist fire, the tropical
forest does not return. The structure of the forest changes and is replaced by pioneer
species, or it may become a savanna (Nobre et al. 2016). This change in structure
leads to forests that are more flammable. That is, if a second fire occurs, the area will
burn at higher intensity, further hindering tree growth.
The Amazon forest is responsible for assimilating a large part of CO2 emissions.
A savannization of the Amazon, due to logging or fire, would thus lead to wide-
spread losses in biodiversity, but also to an increase rate of climate change (Nobre
et al. 2016).
The capacity to assimilate CO2 could also be at stakes in boreal forests. This is
because, like in the Amazon, forest structure also changes after fire and transforms
into grasslands or open woodlands with smaller capacity for absorbing carbon
(Lenton et al. 2008).
Additionally, the current fire suppression policy, where fires are effectively extin-
guished as quickly as possible, has reduced global burned area, as previously men-
tioned. It is estimated that this has enhanced the capacity of C assimilation in land
ecosystems by 19%. In other words, current emissions from fire would increase
from 1.5 to 2.2 GtC year−1 if the burned area was the same as that occurring before
the rural exodus, between 1850 and 1930 (Arora and Melton 2018).

3.1.2.2 Fire Reduces Carbon Stocks

Ground fires may consume large amounts of carbon belowground in cold and tropi-
cal peatlands, which contain vast amounts of carbon. Wildfires in these areas could
potentially unlock the 600–700 GtC that are currently estimated to be stored in these
ecosystems (Turetsky et al. 2015). Examples of such fires are those occurring in
Indonesia in 1997, where ~ 1 GtC was emitted, which was equivalent to 15% of all
fossil fuel emissions at the time (Page et al. 2002).
36 3 Fire as an Earth System Process

Peat fires are naturally rare, because of the high water content in these ecosys-
tems. However, artificial drainage for changing the land cover and use, along with
climate warming-induced drought, lowers the water table and, consequently,
increases the vulnerability of these ecosystems to fire. Burning of these systems has
an effect similar to that of burning fossil fuels. That is, we would be returning to the
atmosphere CO2 that had been removed from the atmosphere for some centuries or
millennia.

3.1.2.3 Fire Increases Recalcitrant Carbon Storage

An additional aspect in fire-climate feedbacks is related to soot production. Soot is


very recalcitrant and difficult to degrade. Consequently, the fraction of organic mat-
ter that is transformed into pyrogenic carbon after a fire may be considered to
become in a long-term storage pool, as it may take centuries or millennia to degrade.
The amount of biomass carbon annually converted into pyrogenic carbon is cur-
rently estimated to be 0.26 Gt C, which is equivalent to 17% of all emissions from
wildfires (Jones et al. 2019).

3.1.2.4 Fire Effects over the Energy Balance

The effects of fire on the energy balance depend on the severity of the burn and on
how the physiognomy of the landscape changes, as well as the biome. In boreal
ecosystems, for instance, fire-induced deforestation cools the climate. This is
because a homogeneous layer of snow would cover the landscape during much of
the year in the absence of trees. The snow reflects back incoming radiation, which
cools the climate (Chapin et al. 2008).
In tropical regions, we encounter the opposite effect, where fire-induced defores-
tation warms the climate (Chapin et al. 2008). This is because evaporative cooling
is very intense in the tropics. That is, much of the incoming radiation is used to
evaporate water, which cools the temperature. Without trees, however, a large por-
tion of the radiation is transformed into sensible heat, which increases the tempera-
ture. Temperate regions show somewhat intermediate effects between boreal and
tropical forests.
Some studies estimate that deforestation could increase surface temperature by
as much as 1.5 °C in tropical and arid ecosystems and by 0.8 °C in temperate
­ecosystems. In contrast, deforestation could decrease minimum temperature by as
much as 0.5 °C in boreal ecosystems (Alkama and Cescatti 2016). Unfortunately,
these estimates do not separate the effects of fire from other agents inducing defor-
estation, and consequently, we are still lacking a direct quantification of fire-induced
biogeophysical feedbacks.
Apart from altering landscape physiognomy, fire also effects albedo, by increas-
ing aerosol depth. This has additional climatic effects. On the one hand, increasing
aerosol concentration has a small cooling effect, as radiation is absorbed by the
3.1 The Earth as a System and the Role of Fire 37

aerosols and does not reach the surface. Additionally, aerosols alter the size of the
condensation nuclei. That is, higher aerosol concentrations lead to smaller cloud
particles which subsequently lead to reduced precipitation (Houghton 2012). There
have been some efforts to quantify these processes, but estimations vary widely
across models. For instance, the values on how aerosols alter the radiative forcing
(i.e., the intensity of cooling effect in this case) vary from −1.1 to −0.1 W m−2
(Hamilton et al. 2018) across models.

3.1.2.5 Fire Effects over the Water Cycle

Wildfires also exert major modifications over the water balance. The initial reduc-
tion in vegetation leads to an increase in runoff. This is because the decrease in
vegetation cover diminishes interception and evapotranspiration, and consequently,
water infiltration quickly reaches field capacity and runoff increases.
The effects of fire on evapotranspiration depend on the severity of the fire as well
as on the post-fire regeneration structure of the community. Evapotranspiration
declined by 41% in a eucalypt resprouting forest from southern Australia that
burned under high severity, relative to an unburned stand. However, evapotranspira-
tion only declined by 3% in the stands burned at moderate severity immediately
after the fire, and evapotranspiration increased by 9% in the 2–3 years post-fire
(Nolan et al. 2014a). The difference in the response between the high- and medium-­
severity sites is likely driven by the difference in leaf area that is consumed during
the fire (higher leaf area implies higher transpiration). The difference between the
stands burned at moderate severity and unburned is likely affected by compensatory
growth (see Chap. 6) and leaf level responses in surviving trees. That is, post-fire
resprouting trees often grow faster than trees in unburned stands, and in addition,
leaf level transpiration is often higher (Nolan et al. 2014b). Additionally, the time to
recover pre-fire evapotranspiration rates was ~10 years. These values are much
lower than in seeder-dominated stands, which could, for instance, be over 100 years
in Eucalyptus regnans stands (Nolan et al. 2015).
How the change in evapotranspiration leads to changes in water catchment is less
well understood due to an overall lack of catchment level data. Overall, water yield
increases but water quality decreases after the fire. Changes in water quality are
driven by fire severity, post-fire plant cover, and soil organic matter (Rust et al.
2019). Organic nitrogen, nitrate, total phosphorous, and total metal concentrations
increase with fire severity, and the effect is less pronounced in areas with high soil
organic matter before the fire. The effects of fire on water quality further vary with
the different ecosystem types and species traits (Harper et al. 2019). That is, whereas
the chemical characterization of ash is nearly universal, the concentration, solubil-
ity, and reactivity of the different elements that compose ash vary markedly with
plant and ecosystem types.
Plant roots as well as soil organic matter stabilize the soil, and their loss during the
fire increases soil erosion. The melting of the organic matter may change the proper-
ties of the organic matter to create a hydrophobic layer, particularly over clay tex-
38 3 Fire as an Earth System Process

tures. Such hydrophobic layer further enhances soil erosion, and the overall process
of fire-induced changes in soil erosion is reasonably well understood (Certini 2005;
Shakesby and Doerr 2006; Shakesby 2011). The change in water quality is indeed a
consequence of the increased erosion. In turn, erosion exerts further climatic effects,
on the one hand, because the transfer of carbon through sediments between the land
and the stream exerts important biogeochemical feedbacks. On the other hand, there
are important nutrient losses during this process, although volatilization during com-
bustion often leads to larger nutrient losses (Dannenmann et al. 2018).

3.2 Fire Before Humans

Fire has played a role in shaping our biosphere for a very long time (Fig. 3.4).
However, our knowledge on how fire has shaped the geological history is relatively
recent. It was not until 1958 (Harris 1958) that charcoal remains were observed for
the first time in the fossil record. Although it had been hypothesized much earlier
that fires could be present in the geological record, interest in that hypothesis “has
died [by 1958] because of lack of fresh evidence” (Harris 1958). The idea that wild-
fires could be important in shaping the Earth system has thus been explored only in
the last few decades. This field has advanced considerably, shaking some long-term
standing assumptions on a negligible role of fire on Earth, which considered fires as
a perturbation of recent origin.
Fire activity in the fossil record has likely been underreported due to difficulties
in recognizing charcoal as an indicator of landscape fires. Significant advancements
in this area have been made in the last two decades, in order to establish the proto-
cols to properly identify charcoal as a fire proxy (Scott et al. 2014; Whitlock and
Larsen 2001). Some uncertainties still remain, however, on the validity of such fos-
sil record to infer fire activity in deep time. Fire activity has been reconstructed from
charcoal (inertinite) remains in the sedimentary record near wetland habitats.
Consequently, most (but not all) fire activity recorded in the fossil will be from sites
near such water bodies, which might skew the interpretation of the fire activity over
geological time scales. This is because the fire regime near wetland habitats might
be very different from that occurring upland. These results should be interpreted
with caution, although similar caveats should be made in vegetation reconstructions
from the sedimentary record.

3.2.1 Early Fires in the Paleozoic

Fires have been on Earth for as long as there have been plants on land (Fig. 3.4). At
the most basic level, a fire requires an ignition source, a fuel, and an oxidizing agent.
Ignition sources have been present throughout Earth’s history as lightning or volca-
nic eruptions. Oxygen levels may have been high enough to support wildfires since
3.2 Fire Before Humans 39

Fig. 3.4 Geological time chart along with variations in oxygen concentration and fire activity,
relative to their maximum value during the Phanerozoic. Data on oxygen concentration and fire
activity from Glasspool and Scott (2010)

the beginning of the Paleozoic era (540 million years ago, mya) (Pausas and Keeley
2009). However, fires could not occur until enough organic matter (i.e., fuel) accu-
mulated on land to carry a fire.
The first evidence of fire on Earth follows briefly the earliest appearance of vas-
cular plants, at the boundary between the Silurian and the Devonian, ~420 mya
(Glasspool et al. 2006). However, wildfires are thought to be rare during that period
because O2 concentrations would have been low during much of Devonian and also
because vegetation was of short stature. The first forests appeared at the mid-­
Devonian (375 mya), and the first widespread occurrence of fire is currently consid-
ered to occur in the Carboniferous ~350 mya (Berner et al. 2003). High fire presence
continued increasing during the Permian and until the end of the Paleozoic.
40 3 Fire as an Earth System Process

The first documented catastrophic wildfire with large-scale impacts, which cas-
caded to affect other trophic levels and nearby ecosystems, dates back from the
Carboniferous in North Mayo, Ireland. A large wildfire at that time is considered to
have triggered large sedimentary deposits into nearby estuaries, increasing water
turbidity and nutrient flushing, consequently killing fishes and other animals due to
anoxia and eutrophication (Falcon-Lang 1998).
The evidence from such a large wildfire activity during the Carboniferous and,
specially, the Permian comes from the large abundances of charcoal. The abundance
of charcoal in modern peats is about 4% in volume, but it may be higher than 20%
in Carboniferous deposits (Scott et al. 2014).
Not only wildfires were present during the Paleozoic, but also fire regimes simi-
lar to those of today seem to have been already present during that time (Pausas and
Keeley 2009). Falcon-Lang (2000) demonstrated the existence of both high-­
frequency low-intensity surface fires and low-frequency high-intensity crown fires
during the Carboniferous. He claimed that savanna-like progymnosperm communi-
ties burned regularly (with a 3–35 years rotation) within the tropical Laurasia, simi-
lar to current subtropical savannas. Furthermore, he also proposed rare fires (with a
rotation of 105–1085 years) in moister Lepidodendron rainforests, which could be
similar to current crown fires in high-latitude conifer forests (Pausas and
Keeley 2009).

3.2.2 Wildfires in the Mesozoic

Fire activity declined with the formation of Pangea at the transition between the
Permian and Triassic, 250 mya (Glasspool and Scott 2010). The formation of the
supercontinent Pangea altered oceanic currents and, consequently, the global cli-
mate. There was a large release of methane, which further increased temperature
and the glaciers retreated. There were also volcanic eruptions, which enhanced acid
rain from sulfur emissions, and it is estimated that temperature rose by up to 8 °C. It
is assumed that up to 90% of all species died during this period, which constitutes
the largest mass extinction event to date. Models suggest O2 levels during this time
dropped down to 15%, the lowest in the Phanerozoic.
It seems likely that conditions for life remained unfavorable for some million
years after the event. However, O2 concentrations and fire systems reestablished by
the end of the Triassic, where O2 could have been above present atmospheric level
(PAL). The boundary between the Triassic and Jurassic is another moment with
mass extinctions, and fire activity is considered to be important during that time.
Some of the fire regimes documented during the Jurassic are also reminiscent of
current fire regimes. For instance, Francis (1984) documented low-intensity surface
fires burning the understory of dry conifer forests in the Upper Jurassic, which
would be similar to the wildfires documented currently for montane conifer species
(Touchan et al. 2012).
3.3 Fire in the Quaternary 41

Another period with important fire activity is the Cretaceous (150–65 mya
(Brown et al. 2012)). Landscapes experienced a large transformation during this
period, where a transition from a gymnosperm- to an angiosperm-dominated world
occurred. It is considered that angiosperms during this time contributed to intensify
a fire regime as they increase fuel loads and favor the development of ladder fuels.
Oxygen concentrations are also considered to be above PAL.
The boundary between the Cretaceous and the Tertiary (~66 mya) is the moment
of the extinction of the dinosaurs. It is now accepted that such extinction event was
induced by the impact of an asteroid, which subsequently induced important
changes in atmospheric chemistry. However, large accumulations of soot were
found in some marine sediments, which led to the hypothesis that such extinction
event would have been driven by a global, asteroid-induced firestorm. Such hypoth-
esis is these days discredited as atmospheric O2 concentrations at the time do not
seem to be high enough to support such a large global firestorm (Scott et al. 2014).
The transition between the late Miocene and the early Pliocene (8–3 mya) saw
the development of the savanna biome, an event that is considered crucial for the
evolution of bipedalism in humans and also an event that is considered to have been
triggered by a surge in fire activity. In a conceptual model using systems analyses
(Beerling and Osborne 2006), it was proposed that an increase in fire activity during
that period favored tree mortality and C4 grass expansion. Concomitantly, increases
in fire activity would have enhanced drought, further declining tree expansion and
enhanced fire activity in a positive self-reinforcing feedback. This hypothesis has
been met with some skepticism (Bond 2014), and alternative views on the origin of
the savanna biome exist (Charles-Dominique et al. 2016). Other authors have argued
that the increase in fire activity may have been driven by the explosion of a super-
nova, which would have increased atmospheric ionization and, consequently,
increased lightning activity (Melott and Thomas 2019).

3.3 Fire in the Quaternary

3.3.1  umans Increased Burned Area in the Pleistocene


H
and Decreased It During the Holocene in Africa

The first species from the genus Homo (humans) appeared on Earth some 2.5 mya
in Africa. It took them some time to dominate fire but, once they did, the Earth’s
biosphere and climate as well as the evolution of hominins would be radically modi-
fied. Archibald et al. (2012) proposed that there are six stages in the evolution of the
fire-hominins relationship. There are a series of assumption underlying this analy-
sis, and the actual and dating and magnitude of the effects might not be correct.
However, it serves as a first approximation towards understanding the affects of
human activities on burned area.
42 3 Fire as an Earth System Process

In the first stage, lightning was the only ignition source, and hominins lacked the
capacity to use such fires. In the second stage, hominins had learned the ability to
use lightning fires. This occurred at least 1 mya (Berna et al. 2012), although indi-
rect evidences point towards an earlier fire use, extending possibly up to 1.9 mya. It
is established that hominin started using fire for cooking during this period. However,
it is much more difficult to establish whether the landscape was being altered by
hominin fire use, as separating natural from hominin-induced fire use in the char-
coal record is challenging. An adventurous assumption is that, if hominins had the
capacity to create fire landscapes by igniting grassing fuels, then they used such
capacity and could have increased the frequency of fire events. If this assumption is
true, we could expect a modest increase in burned area as a result.
In the third stage, hominins learned to ignite their own fires, which is thought to
occur between 200 kya and 400 kya (thousand years ago) (Clark and Harris 1985).
Lightning-induced fires in the African savanna occur during thunderstorms in the
wet season. However, the capacity to control fires would have allowed hominins to
start fires during the dry season for some agricultural practices. This would have led
to another modest increase in burned area. The fourth stage is characterized by an
increase in human population, occurring some 70–4 kya (Mellars 2006), which sub-
sequently dispersed and leads to further increases in burned area.
In the fifth stage, animal domestication and land cultivation in Africa result in
large fuel discontinuities that, consequently, decrease fire activity, starting 4 kya
(Roberts 1998). In the sixth stage, starting 200 cal. year BP, there is a fast and large
population growth, which leads to further increases in fuel discontinuity and conse-
quently leads to further declines in fire activity.
It is worth noting that this view of early human domination of fire regimes, and
of large-scale landscape transformations, is currently being disputed. Some studies,
for instance, show that climatic influences are larger than human effects in South
Africa for the past 170,000 years (Daniau et al. 2013). In Europe, additionally, fire
regimes were not immediately disrupted after human colonization. On the contrary,
fire activity during the Holocene followed climatic oscillations, and there is little
evidence of large-scale disruptions on the fire regime until much later, once human
populations were much larger (Lawson et al. 2013; Daniau et al. 2010).

3.3.2  uman Alterations to the Fire Regime During


H
the Holocene

The previous example from the African savannas serves to exemplify that fire use by
hominins has accompanied human evolution and development for a long period of
time. There are countless examples of how humans have modified landscapes during
the Holocene for agricultural activities. Broadly speaking, these may be considered
as fire-stick farming, slash and burn farming, and fire-forage pastoralism. The rela-
tive importance of these practices as drivers of global changes in the fire regime is
3.3 Fire in the Quaternary 43

still discussed. For instance, some global synthesis studies highlight how human
activities may have locally altered fire regimes, but large-scale patterns are often bet-
ter explained after including interactions with climate variations (Marlon et al. 2013).

3.3.2.1 Fire-Stick Farming

Fire-stick farming refers to the set of practices that, altogether, use fire as a tool for
changing the type of land cover and use, for hunting, killing vermin, or creating new
travel paths, among others. Fire-stick farming has deep consequences for ecosystem
functioning as well as for biosphere-atmosphere interactions. For instance, early
human colonizers that arrived into Australia ~45,000 cal. year BP used fire as a tool
to transform the landscape from closed canopy vegetation into a variety of habitats
that included pastures or open woodlands. Such landscape transformation was
accompanied by the extinction of larger browsing mammals (van der Kaars et al.
2017). Furthermore, such land clearing could have increased aridification over the
country because evapotranspiration was reduced and albedo increased. Some
authors have described this phenomenon as an ecosystem “collapsing” in Australian
land (Miller et al. 2005). It should be noted that this hypothesis is not exempt from
criticism, as other authors view climate change, rather than humans, as responsible
for the drivers of Australian fire regimes (Mooney et al. 2011).
The Australian case shows how ecosystems recover stability and how novel eco-
systems emerge by moving to a new state after a large-scale perturbation on the fire
regime. Indeed, despite the dramatic consequences of initial human colonization,
extensive and regular human fire use over 40,000 years had created a variety of
“pyrohabitats” upon which the biota was reliant (Scott et al. 2014). The arrival of
European colonizers ~200 cal. year BP provided a new disruption to the traditional
fire management in the area (Mooney et al. 2011), which allowed for larger fuel
build-ups and, consequently, less abundant but more catastrophic wildfires.
It should be noted that the Australian case is not unique and, in fact, similar sto-
ries, where humans have transformed landscape physiognomy, which exerts cascad-
ing effects over trophic levels, have occurred repetitively throughout human history.
Human colonization in the Americas and the use of fire, for instance, led to the
extinction of the large mammals around 14,000 cal. year BP (Gill et al. 2009). The
removal of large herbivores, in turn, would have further exacerbated fire activity.
That is, fires took over the role of vegetation consumer in the absence of herbivores.
In Spain, the expansion of agricultural areas since the Neolithic led to recurrent
land clearing through fire, which also exerted large-scale physiognomic changes
(Valbuena-Carabaña et al. 2010). Many conifer species, which had shown resilience
to climatic changes and to the natural fire regime during the Holocene, were not able
to survive under the new, more recurrent and intense fire regime imposed by human
settlements, and they were displaced by resprouter species, such as oaks (Morales-­
Molino et al. 2017) (Fig. 3.5). Such changes in land cover have repeated over this
landscape, further favoring the replacement of resprouter over non-resprouting tree
44 3 Fire as an Earth System Process

Fig. 3.5 Example of fire use for land use change. X-axis represents age in cal. year BP. The tri-
angles indicate periods with maximum fire activity; the three on the right were linked to dry phases,
but the ones at the left were not. Shaded areas indicate, from right to left, Iron Age settlements,
Roman settlements, and Christian-Muslims wars. Pinus had dominated throughout much of the
Holocene in this landscape from Central Spain, and it had been able to survive under the natural
fire regime, including the peaks in fire activity between 6000 and 4000 cal. year BP. Fire activity
declines after grazing activities increase (indicated by increases in coprophilous fungi) during
3000–4000 cal. year BP. The highly recurrent presence of high-intensity and high-severity fires
experienced around 1400–1200 cal. year BP led to the demise of pine in the area, which was sub-
sequently used for agricultural activities. (Reproduced with permission from Morales-Molino
et al. (2017))

species, during the Copper, Bronze, and Iron Ages, in a practice that continued well
into the nineteenth century. The only exception has been conifers with aerial seed
banks, which are adapted to crown fire regimes and were able survive the anthropo-
genic fire regime of higher fire frequency.

3.3.2.2 Slash and Burn Farming

Slash and burn has been a traditional and widespread form of land management that
exerts important feedbacks on the Earth system. This practice consists of letting the
bush grow for 10–20 years, after which the shrubs are slashed, piled, and burned.
The ashes are then used as a fertilizer over the cultivation area, where crops are
grown for a period of about 5 years, after which the land is let to lie fallow and
restart the cycle. This practice is common throughout much of Africa, Asia, and
Latin America and exerts important climatic feedbacks. For instance, the area under
slash and burn in America was drastically reduced after European colonization
(Koch et al. 2019). Diseases introduced by colonizers decimated the populations of
Indians, and consequently, fire activity in the Americas declined during the 1500s.
Such a decrease in fire activity, along with reforestation and other bioclimatic feed-
backs, led to an overall drawdown of atmospheric CO2 of ~ 5 ppm, which might
have contributed to the cooling of ~0.15 °C observed during that time (Koch
et al. 2019).
3.4 Fire in the Anthropocene 45

Slash and burn is also responsible for a large percentage of the carbon emissions
that currently originate from fire. Fire and land use change currently emit 1.5 GtC
yr-1, which is equivalent to 14% of total greenhouse gas emissions (Le Quéré et al.
2018). Seventy percent of these emissions originate in the savanna biome (Giglio
et al. 2013), where slash and burn is most common. Furthermore, there is a worrying
drying trend in the Congo Basin, an area that corresponds to the second largest tract
of tropical forest in the world (after the Amazon) and is surrounded by savannas. The
fire in the region could be a contributing factor to such drying trend as the smoke
from fire declines precipitation in the area by inhibiting convection (Tosca et al. 2015).

3.3.2.3 Fire-Forage Pastoralism

Fire rejuvenates landscapes. This has been the rationale for using fire as a tool for
pastoralism, where rangelands are regularly burned for developing fresh pasture.
The problem is that fire also degrades landscapes, particularly over steep slopes and
in erosion-prone areas. One example comes from transhumance, a common practice
across Europe, where flocks are transported to different areas in different seasons.
Fires were regularly used to clear the land and grazers fed on the newly emerged
grass. In fact, some authors claim that the etymological origin of transhumance
comes from merging the Latin trans (across) with fumo (smoke), indicating that
flocks were transported to smoking (burned) forests (Valbuena-Carabaña et al.
2010). Over time, this led to the degradation over, for instance, large areas in the
Iberian Peninsula, where lands that once were covered by forests are now barren
lands, with thin and unfertile soils (Valbuena-Carabaña et al. 2010).
Fire can also be combined with land cultivation, such that extensive and intensive
farming can be combined. That is, fires are used for rejuvenating the grass, with or
without land cultivation, for extensive grazing, and crops provide food for the times
when animals are in the stable.

3.4 Fire in the Anthropocene

3.4.1 Changes in the Fire Regime from Recent Human Activity

Industrialization and large population increases have created in recent times, once
again, another modification of the fire regimes. We previously mentioned (Chap. 2)
the existence of a pyrome common across all biomes (Fig. 2.5), likely resulting
from human activity. Such Human activity has created a fire regime where fires are
usually small because they are put out as soon as it is possible. This “fire suppres-
sion” policy has led to the increase in catastrophic wildfires (Biswell 1999). This is
because fire in many areas would naturally decline fuel availability, but absence of
fire leads to large accumulations which then burn in high-intensity wildfires.
46 3 Fire as an Earth System Process

Furthermore, rural areas in Europe have experienced an exodus in large areas of the
world, which further contributes to increasing fuel loads. Thus, our highly efficient
firefighting systems have been able to reduce overall fire activity so far. The ques-
tion is for how long will such a decline in burned area be maintained.
Some of the mega-fires experienced in recent years, sometimes called sixth-­
generation fires, have burned at unprecedented rates, as we will discuss in Chap. 9.
Additional factors contributing to the rise of mega-fires include the expansion of
reforestations, which were subsequently abandoned, in S America and the
Mediterranean Basin (Gómez-González et al. 2018) and the spread of invasive spe-
cies in N America, S America, and S Africa (Fusco et al. 2019; Moreira et al. 2019).
Poor urbanization planning in N America and Australia (Penman et al. 2016;
Radeloff et al. 2018) further creates problems at the wildland urban interface (see
Chap. 1). Overall, the influence of human activities over wildfire activity is increas-
ing in many areas worldwide as a direct result of human activities (Syphard et al.
2017; Turco et al. 2016) and as an indirect result of anthropogenic climate change
(Abatzoglou and Williams 2016; Boer et al. 2020). We will develop to a greater
extent the human alteration of fire regimes, as well as how to tackle fire manage-
ment, in Chap. 9.

3.4.2 Wildfires and Health

It is worth noting that wildfires also influence human health. Wildfires release large
quantities of smoke to the atmosphere. Smoke is mainly composed of particulate
matter and that with a diameter of less than 2.5 μm (PM2.5) is particularly problem-
atic. Large concentrations of such PM2.5 are toxic, and they have been associated
with increased neonatal and cardiorespiratory mortality, worsening of respiratory
and cardiovascular problems, and physiological changes including inflammation,
oxidative stress, and procoagulation (Bowman and Johnston 2005; Kunzli et al.
2006; Johnston et al. 2012). 339,000 deaths annually have been attributed to smoke
derived from landscape fires (Johnston et al. 2012).
Smoke-induced fatalities are mostly localized in sub-Saharan Africa (157,000
deaths annually), the area where fires are most common globally (Table 1.1), fol-
lowed by Southeast Asia (110,000 deaths annually, Fig. 3.6). PM emissions depend
on the fire type and also on the vegetation consumed (Hu et al. 2018). Ground fires
consistently emit higher amounts of PM than surface or crown fires because smol-
dering combustion, which dominates in ground fires, releases more PM (Hu et al.
2018). Peat fires dominate in Southeast Asia, which could help explain the very high
mortality rates after fire in the area.
The vegetation that is consumed during the fire may also affect the amount of PM
emitted. The emission factor (EF), which indicates the amount of an ion emitted per
unit of fuel consumed, often varies between conifers and broadleaved species and
also between trees and grasses. This implies that fires in some vegetation types will
show more adverse effects on health than others (Ma et al. 2019; Urbanski et al. 2008).
References 47

Fig. 3.6 Annual number deaths attributed to smoke from landscape fires. (Reproduced from
Johnston et al. (2012))

References

Abatzoglou JT, Williams AP (2016) Impact of anthropogenic climate change on wildfire across
western US forests. Proc Natl Acad Sci 113(42):11770–11775. https://doi.org/10.1073/
pnas.1607171113
Alkama R, Cescatti A (2016) Biophysical climate impacts of recent changes in global forest cover.
Science 351(6273):600–604. https://doi.org/10.1126/science.aac8083
Archibald S, Staver AC, Levin SA (2012) Evolution of human-driven fire regimes in Africa. Proc
Natl Acad Sci U S A 109(3):847–852. https://doi.org/10.1073/pnas.1118648109
Arora VK, Melton JR (2018) Reduction in global area burned and wildfire emissions since
1930s enhances carbon uptake by land. Nat Commun 9(1):1326. https://doi.org/10.1038/
s41467-018-03838-0
Beerling DJ, Osborne CP (2006) The origin of the savanna biome. Glob Chang Biol 12:2023–2031
Belcher CM, Yearsley JM, Hadden RM, McElwain JC, Rein G (2010) Baseline intrinsic flamma-
bility of Earth’s ecosystems estimated from paleoatmospheric oxygen over the past 350 million
years. Proc Natl Acad Sci 107(52):22448–22453. https://doi.org/10.1073/pnas.1011974107
Berna F, Goldberg P, Horwitz LK, Brink J, Holt S, Bamford M, Chazan M (2012) Microstratigraphic
evidence of in situ fire in the Acheulean strata of Wonderwerk Cave, Northern Cape prov-
ince, South Africa. Proc Natl Acad Sci U S A 109(20):E1215–E1220. https://doi.org/10.1073/
pnas.1117620109
Berner RA, Beerling DJ, Dudley R, Robinson JM, Wildman RA (2003) Phanerozoic atmo-
spheric oxygen. Annu Rev Earth Planet Sci 31(1):105–134. https://doi.org/10.1146/annurev.
earth.31.100901.141329
Bierregaard RO, Gascon C, Lovejoy TE, Mesquita R (2001) Lessons from Amazonia – the ecology
and conservation of a fragmented forest. Yale University Press, Yale
Biswell H (1999) Prescribed burning in California Wildlands vegetation management. University
of California Press, Los Angeles
Boer MM, Resco de Dios V, Bradstock RA (2020) Unprecedented burn area of Australian mega
forest fires. Nat Clim Chang 10:171–172
Bond WJ (2014) Fires in the Cenozoic: a late flowering of flammable ecosystems. Front Plant Sci
5:749. https://doi.org/10.3389/fpls.2014.00749
Bowman DMJS, Johnston FH (2005) Wildfire smoke, fire management, and human health.
EcoHealth 2(1):76–80. https://doi.org/10.1007/s10393-004-0149-8
Brown SAE, Scott AC, Glasspool IJ, Collinson ME (2012) Cretaceous wildfires and their impact
on the Earth system. Cretac Res 36:162–190. https://doi.org/10.1016/j.cretres.2012.02.008
48 3 Fire as an Earth System Process

Certini G (2005) Effects of fire on properties of forest soils: a review. Oecologia 143(1):1–10.
https://doi.org/10.1007/s00442-004-1788-8
Chapin FS, Randerson JT, McGuire AD, Foley JA, Field CB (2008) Changing feedbacks in the
climate–biosphere system. Front Ecol Environ 6(6):313–320. https://doi.org/10.1890/080005
Charles-Dominique T, Davies J, Hempson G, Simmy B, Daru B, Kabongo R, Maurin O, Muasya
A, Bank M, Bond W (2016) Spiny plants, mammal browsers, and the origin of African savan-
nas. Proc Nat Acad Sci 113. https://doi.org/10.1073/pnas.1607493113
Clark JD, Harris JWK (1985) Fire and its roles in early hominid lifeways. Afr Achael Rev 3:3–27
Cochrane MA, Schulze MD (1998) Forest fires in the Brazilian Amazon. Conserv Biol
12(5):948–950
Daniau AL, d’Errico F, Sanchez Goni MF (2010) Testing the hypothesis of fire use for ecosystem
management by neanderthal and upper palaeolithic modern human populations. PLoS One
5(2):e9157. https://doi.org/10.1371/journal.pone.0009157
Daniau AL, Sanchez Goni MF, Martinez P, Urrego DH, Bout-Roumazeilles V, Desprat S, Marlon
JR (2013) Orbital-scale climate forcing of grassland burning in southern Africa. Proc Natl
Acad Sci U S A 110(13):5069–5073. https://doi.org/10.1073/pnas.1214292110
Dannenmann M, Díaz-Pinés E, Kitzler B, Karhu K, Tejedor J, Ambus P, Parra A, Sánchez L, Resco
V, Ramírez D, Povoas-Guimaraes L, Zechmeister-Boltenstern S, Kraus D, Castaldi S, Vallejo
A, Rubio A, Moreno J, Butterbach-Bahl K (2018) Post-fire nitrogen balance of Mediterranean
shrublands: direct combustion losses versus gaseous and leaching losses from the post-fire soil
mineral nitrogen flush. Glob Chan Biol 24(10):4505–4520
Falcon-Lang H (1998) The impact of wildfire on an early carboniferous coastal environment,
North Mayo, Ireland. Palaeogeogr Palaeoclimatol Palaeoecol 139:121–138
Falcon-Lang HJ (2000) Fire ecology of the Carboniferous tropical zone. Palaeogeogr Palaeoclimatol
Palaeoecol 164:339–355
Francis JE (1984) The seasonal environment of the Purbeck (Upper Jurassic) fossil forests.
Palaeogeogr Palaeoclimatol Palaeoecol 48:285–307
Friedlingstein P, Meinshausen M, Arora VK, Jones CD, Anav A, Liddicoat SK, Knutti R (2014)
Uncertainties in CMIP5 climate projections due to carbon cycle feedbacks. J Clim 27:511–526.
https://doi.org/10.1175/jcli-d-12-00579.1
Fusco EJ, Finn JT, Balch JK, Nagy RC, Bradley BA (2019) Invasive grasses increase fire occur-
rence and frequency across US ecoregions. Proc Natl Acad Sci 116(47):23594–23599. https://
doi.org/10.1073/pnas.1908253116
Giglio L, Randerson JT, van der Werf GR (2013) Analysis of daily, monthly, and annual burned
area using the fourth-generation global fire emissions database (GFED4). J Geophys Res
Biogeo 118:317–328. https://doi.org/10.1002/jgrg.20042
Gil-Romera G, González-Sampériz P, Lasheras-Álvarez L, Sevilla-Callejo M, Moreno A, Valero-­
Garcés B, López-Merino L, Carrión JS, Pérez Sanz A, Aranbarri J, García-Prieto Fronce E
(2014) Biomass-modulated fire dynamics during the last glacial–interglacial transition at
the Central Pyrenees (Spain). Palaeogeogr, Palaeoclim, Palaeoecol 402:113–124. https://doi.
org/10.1016/j.palaeo.2014.03.015
Gill JL, Williams JW, Jackson ST, Lininger KB, Robinson GS (2009) Pleistocene Megafaunal
collapse, novel plant communities, and enhanced fire regimes in North America. Science
326(5956):1100–1103. https://doi.org/10.1126/science.1179504
Glasspool IJ, Edwards D, Axe L (2006) Charcoal in the early Devonian: a wildfire-derived
Konservat–Lagerstätte. Rev Palaeobot Palynol 142(3):131–136. https://doi.org/10.1016/j.
revpalbo.2006.03.021
Glasspool IJ, Scott AC (2010) Phanerozoic concentrations of atmospheric oxygen reconstructed
from sedimentary charcoal. Nat Geosci 3:627–630. https://doi.org/10.1038/ngeo9231038/
NGEO923
References 49

Gómez-González S, Ojeda F, Fernandes PM (2018) Portugal and Chile: longing for sustainable
forestry while rising from the ashes. Environ Sci Pol 81:104–107. https://doi.org/10.1016/j.
envsci.2017.11.006
Hamilton DS, Hantson S, Scott CE, Kaplan JO, Pringle KJ, Nieradzik LP, Rap A, Folberth
GA, Spracklen DV, Carslaw KS (2018) Reassessment of pre-industrial fire emissions
strongly affects anthropogenic aerosol forcing. Nat Commun 9(1). https://doi.org/10.1038/
s41467-018-05592-9
Harper AR, Santin C, Doerr SH, Froyd CA, Albini D, Otero XL, Viñas L, Pérez-Fernández B
(2019) Chemical composition of wildfire ash produced in contrasting ecosystems and its toxic-
ity to Daphnia magna. Int J Wildland Fire 28(10):726. https://doi.org/10.1071/wf18200
Harris TM (1958) Forest fire in the Mesozoic. J Ecol 46:447–453
Houghton J (2012) Global warming – the complete briefing. Cambridge University Press,
Cambridge
Hu Y, Fernandez-Anez N, Smith TEL, Rein G (2018) Review of emissions from smouldering peat
fires and their contribution to regional haze episodes. Int J Wildland Fire 27(5):293. https://doi.
org/10.1071/wf17084
Johnston FH, Henderson SB, Chen Y, Randerson JT, Marlier M, Defries RS, Kinney P, Bowman
DM, Brauer M (2012) Estimated global mortality attributable to smoke from landscape fires.
Environ Health Perspect 120(5):695–701. https://doi.org/10.1289/ehp.1104422
Jones MW, Santín C, van der Werf GR, Doerr SH (2019) Global fire emissions buffered by
the production of pyrogenic carbon. Nat Geosci 12(9):742–747. https://doi.org/10.1038/
s41561-019-0403-x
Koch A, Brierley C, Maslin MM, Lewis SL (2019) Earth system impacts of the European arrival
and great dying in the Americas after 1492. Quat Sci Rev 207:13–36. https://doi.org/10.1016/j.
quascirev.2018.12.004
Kump LR (1988) Terrestrial feedback in atmospheric oxygen regulation by fire and phosphorus.
Nature 335(6186):152–154. https://doi.org/10.1038/335152a0
Kunzli N, Avol E, Wu J, Gauderman WJ, Rappaport E, Millstein J, Bennion J, McConnell R,
Gilliland FD, Berhane K, Lurmann F, Winer A, Peters JM (2006) Health effects of the 2003
Southern California wildfires on children. Am J Respir Crit Care Med 174(11):1221–1228.
https://doi.org/10.1164/rccm.200604-519OC
Lawson IT, Tzedakis PC, Roucoux KH, Galanidou N (2013) The anthropogenic influence on wild-
fire regimes: charcoal records from the Holocene and Last Interglacial at Ioannina, Greece. J
Biogeogr 40(12):2324–2334. https://doi.org/10.1111/jbi.12164
Le Quéré C, Andrew RM, Friedlingstein P, Sitch S, Hauck J, Pongratz J, Pickers PA, Korsbakken
JI, Peters GP, Canadell JG, Arneth A, Arora VK, Barbero L, Bastos A, Bopp L, Chevallier F,
Chini LP, Ciais P, Doney SC, Gkritzalis T, Goll DS, Harris I, Haverd V, Hoffman FM, Hoppema
M, Houghton RA, Hurtt G, Ilyina T, Jain AK, Johannessen T, Jones CD, Kato E, Keeling RF,
Goldewijk KK, Landschützer P, Lefèvre N, Lienert S, Liu Z, Lombardozzi D, Metzl N, Munro
DR, Nabel JEMS, S-i N, Neill C, Olsen A, Ono T, Patra P, Peregon A, Peters W, Peylin P,
Pfeil B, Pierrot D, Poulter B, Rehder G, Resplandy L, Robertson E, Rocher M, Rödenbeck
C, Schuster U, Schwinger J, Séférian R, Skjelvan I, Steinhoff T, Sutton A, Tans PP, Tian H,
Tilbrook B, Tubiello FN, van der Laan-Luijkx IT, van der Werf GR, Viovy N, Walker AP,
Wiltshire AJ, Wright R, Zaehle S, Zheng B (2018) Global Carbon Budget 2018. Earth Syst Sci
Data 10(4):2141–2194. https://doi.org/10.5194/essd-10-2141-2018
Lenton T (2016) Earth system science: a very short introduction. Oxford University Press, Oxford
Lenton TM, Held H, Kriegler E, Hall JW, Lucht W, Rahmstorf S, Schellnhuber HJ (2008) Tipping
elements in the Earth’s climate system. Proc Natl Acad Sci 105(6):1786–1793. https://doi.
org/10.1073/pnas.0705414105
Ma Y, Tigabu M, Guo X, Zheng W, Guo L, Guo F (2019) Water-soluble inorganic ions in fine
particulate emission during forest fires in chinese boreal and subtropical forests: an indoor
experiment. Forests 10(11):994. https://doi.org/10.3390/f10110994
50 3 Fire as an Earth System Process

Marlon JR, Bartlein PJ, Daniau A-L, Harrison SP, Maezumi SY, Power MJ, Tinner W, Vanniére B
(2013) Global biomass burning: a synthesis and review of Holocene paleofire records and their
controls. Quat Sci Rev 65:5–25. https://doi.org/10.1016/j.quascirev.2012.11.029
Mellars P (2006) Why did modern human populations disperse from Africa ca. 60,000 years ago?
A new model. Proc Natl Acad Sci U S A 103:9381–9386
Melott AL, Thomas BC (2019) From cosmic explosions to terrestrial fires? J Geol 127:475–481.
https://doi.org/10.1086/703418
Miller GH, Fogel ML, Magee JW, Gagan MK, Clarke SJ, Johnson BJ (2005) Ecosystem collapse
in Pleistocene Australia and a human role in Megafaunal extinction. Science 309(5732):287–
290. https://doi.org/10.1126/science.1111288
Mooney SD, Harrison SP, Bartlein PJ, Daniau AL, Stevenson J, Brownlie KC, Buckman S, Cupper
M, Luly J, Black M, Colhoun E, D’Costa D, Dodson J, Haberle S, Hope GS, Kershaw P,
Kenyon C, McKenzie M, Williams N (2011) Late quaternary fire regimes of Australasia. Quat
Sci Rev 30(1):28–46. https://doi.org/10.1016/j.quascirev.2010.10.010
Morales-Molino C, Tinner W, García-Antón M, Colombaroli D (2017) The historical demise of
Pinus nigra forests in the Northern Iberian Plateau (South-Western Europe). J Ecol 105:634–646
Moreira F, Ascoli D, Safford H, Adams M, Moreno JM, Pereira JC, Catry F, Armesto J, Bond
WJ, Gonzalez M, Curt T, Koutsias N, McCaw L, Price O, Pausas J, Rigolot E, Stephens S,
Tavsanoglu C, Vallejo R, Van Wilgen B, Xanthopoulos G, Fernandes P (2019) Wildfire man-
agement in Mediterranean-type regions: paradigm change needed. Environ Res Lett doi:https://
doi.org/10.1088/1748-9326/ab541e
Nobre CA, Sampaio G, Borma LS, Castilla-Rubio JC, Silva JS, Cardoso M (2016) Land-use and
climate change risks in the Amazon and the need of a novel sustainable development paradigm.
Proc Natl Acad Sci 113(39):10759–10768. https://doi.org/10.1073/pnas.1605516113
Nolan RH, Lane PNJ, Benyon RG, Bradstock RA, Mitchell PJ (2014a) Changes in evapotranspira-
tion following wildfire in resprouting eucalypt forests. Ecohydrology 7:1363–1377. https://doi.
org/10.1002/eco.1463
Nolan RH, Lane PNJ, Benyon RG, Bradstock RA, Mitchell PJ (2015) Trends in evapotranspira-
tion and streamflow following wildfire in resprouting eucalypt forests. J Hydrol 524:614–624.
https://doi.org/10.1016/j.jhydrol.2015.02.045
Nolan RH, Mitchell PJ, Bradstock RA, Lane PN (2014b) Structural adjustments in resprout-
ing trees drive differences in post-fire transpiration. Tree Physiol 34(2):123–136. https://doi.
org/10.1093/treephys/tpt125
Page SE, Siegert F, Rieley JO, Boehm H-DV, Jaya A, Limin S (2002) The amount of carbon
released from peat and forest fires in Indonesia during 1997. Nature 420(6911):61–65. https://
doi.org/10.1038/nature01131
Pausas JG, Keeley JE (2009) A burning story: the role of fire in the history of life. Bioscience
59(7):593–601. https://doi.org/10.1525/bio.2009.59.7.10
Penman TD, Eriksen CE, Horsey B, Bradstock RA (2016) How much does it cost residents to pre-
pare their property for wildfire? Int J Disaster Risk Reduct 16:88–98. https://doi.org/10.1016/j.
ijdrr.2016.01.012
Radeloff VC, Helmers DP, Kramer HA, Mockrin MH, Alexandre PM, Bar-Massada A, Butsic V,
Hawbaker TJ, Sn M, Syphard AD, Stewart SI (2018) Rapid growth of the US wildland-urban
interface raises wildfire risk. Proc Natl Acad Sci U S A 115:3314–3319
Roberts N (1998) The Holocene: an environmental history. Wiley-Blackwell, Oxford
Rust AJ, Saxe S, McCray J, Rhoades CC, Hogue TS (2019) Evaluating the factors responsible for
post-fire water quality response in forests of the western USA. Int J Wildland Fire 28(10):769.
https://doi.org/10.1071/wf18191
Scott AC, Bowman DMJS, Bond WJ, Pyne SJ, Alexander ME (2014) Fire on earth: an introduc-
tion. Wiley-Blackwell, Chichester
Shakesby R, Doerr S (2006) Wildfire as a hydrological and geomorphological agent. Earth Sci Rev
74(3–4):269–307. https://doi.org/10.1016/j.earscirev.2005.10.006
References 51

Shakesby RA (2011) Post-wildfire soil erosion in the Mediterranean: review and future research
directions. Earth Sci Rev 105(3–4):71–100. https://doi.org/10.1016/j.earscirev.2011.01.001
Syphard AD, Keeley JE, Pfaff AH, Ferschweiler K (2017) Human presence diminishes the
importance of climate in driving fire activity across the United States. Proc Natl Acad Sci
114(52):13750–13755. https://doi.org/10.1073/pnas.1713885114
Tosca MG, Diner DJ, Garay MJ, Kalashnikova OV (2015) Human-caused fires limit convection
in tropical Africa: first temporal observations and attribution. Geophys Res Lett 42(15):6492–
6501. https://doi.org/10.1002/2015gl065063
Touchan R, Baisan C, Mitsopoulos ID, Dimitrakopoulos AP (2012) Fire history in European Black
Pine (Pinus nigra Arn.) forests of the Valia Kalda, Pindus Mountains, Greece. Tree Ring Res
68(1):45–50. https://doi.org/10.3959/2011-12.1
Turco M, Bedia J, Di Liberto F, Fiorucci P, von Hardenberg J, Koutsias N, Llasat M-C, Xystrakis
F, Provenzale A (2016) Decreasing fires in Mediterranean Europe. PLoS One 11(3):e0150663.
https://doi.org/10.1371/journal.pone.0150663
Turetsky MR, Benscoter B, Page S, Rein G, GRvd W, Watts A (2015) Global vulnerability of
peatlands to fire and carbon loss. Nat Geosci 8:11–14. https://doi.org/10.1038/ngeo23251038/
NGEO2325
Urbanski SP, Hao WM, Baker S (2008) Chemical composition of wildland fire emissions. In:
Bytnerowicz A, Arbaugh M, Riebau A, Andersen C (eds) Developments in environmental sci-
ence, vol 8. Elsevier, pp 79–107. https://doi.org/10.1016/s1474-8177(08)00004-1
Valbuena-Carabaña M, de Heredia UL, Fuentes-Utrilla P, González-Doncel I, Gil L (2010)
Historical and recent changes in the Spanish forests: a socio-economic process. Rev Palaeobot
Palynol 162(3):492–506. https://doi.org/10.1016/j.revpalbo.2009.11.003
van der Kaars S, Miller GH, Turney CS, Cook EJ, Nurnberg D, Schonfeld J, Kershaw AP, Lehman
SJ (2017) Humans rather than climate the primary cause of Pleistocene megafaunal extinction
in Australia. Nat Commun 8:14142. https://doi.org/10.1038/ncomms14142
Watson A, Lovelock JE (2013) The dependence of flame spread and probability of ignition on
atmospheric oxygen. In: Belcher C (ed) Fire phenomena and the earth system. Wiley, West
Sussex, pp 273–287. https://doi.org/10.1002/9781118529539.ch14
Watson A, Lovelock JE, Margulis L (1978) Methanogenesis, fires and the regulation of atmo-
spheric oxygen. Biosystems 10(4):293–298. https://doi.org/10.1016/0303-2647(78)90012-6
Whitlock C, Larsen C (2001) Charcoal as a fire proxy. In: Smol JP, Birks HJB, Last WM (eds)
Tracking environmental change using lake sediments, Terrestrial, algal, and siliceous indica-
tors, vol 3. Kluwer Academic Publishers, Dordrecht, pp 75–97
Chapter 4
The Evolution of Physiological Adaptations
in a Flammable Planet

Abstract Fire has been present on Earth since vascular plant expansion over land
at the turn of the Devonian, 420 mya. It is thus to be expected that wildfires have
shaped physiological plant traits. The question is which traits have evolved under
fire activity and to which extent. Here we present the general hypothesis that plants
have evolved bet hedging strategies, whenever possible, to jointly deal with the
multiple stresses and disturbances they face during their evolutionary lifetimes. We
also note that, in some instances, trade-offs in the adaptations to stress and distur-
bance have also developed. The traits that are most likely to have evolved in response
to fire are related to post-fire recruitment and persistence. We argue that the evolu-
tion of these traits has also been shaped by drought and other environmental factors.
We also discuss the evidence on whether plant traits that enhance flammability
could have evolved in response to fire, but we find more plausible the explanation
that plant traits affecting fire behavior emerge from plant responses to other envi-
ronmental factors.

4.1 Introduction

We have discussed in the previous chapter how fires have been present across much
of the Earth’s history, affecting processes such as O2 concentration or the Earth’s
climate. Fires have been present on Earth for the past ~400 mya, that is, for nearly
as long as there have been plants on land. It thus seems logical to expect that fire has
been an agent shaping plant physiological evolution, at least to some degree.
However, there are diverging views in the evolutionary literature on the potential
role of fire. Those views range from not mentioning fire as an agent in plant evolu-
tion at all (Niklas 2016) to considering fire as a major driver of plant evolution
(Keeley et al. 2012). There are different reasons for such a diversity of views. For
instance, one of the causes underlying the assumption of a negligible role of fire is
that the importance of fire in the geological record has been underrated (see Chap. 3).
One of the major discussions on the role of fire as a driver of plant evolution is to
understand whether physiological traits that enhance post-fire survival and repro-

© Springer Nature Switzerland AG 2020 53


V. Resco de Dios, Plant-Fire Interactions, Managing Forest Ecosystems 36,
https://doi.org/10.1007/978-3-030-41192-3_4
54 4 The Evolution of Physiological Adaptations in a Flammable Planet

duction are adaptations or exaptations (Lamont and He 2017). That is, some authors
consider that traits enhancing fitness under fire-prone environments actually evolved
in response to another environmental pressure (Bradshaw et al. 2011). For instance,
resprouting could have evolved in response to herbivory or other disturbances. If
true, then resprouting would be an exaptation, not an adaptation, to fire.
Understanding whether plant evolution has been shaped by fire is important to
develop an evolutionary theoretical basis to fire effects. However, discussions on
whether traits enhancing fitness are an adaptation or an exaptation to fire are, at least
to a point, sterile. Plants face multiple challenges throughout their lives, and they
thus need to concomitantly respond to the different environmental stresses and dis-
turbances that may occur (Fig. 4.1). Instead of assuming that a trait has evolved in
response to a single factor, it may be more plausible to assume that it has evolved in
response to multiple environmental factors, at least in some instances. At the very
least, we should acknowledge the possibility that trait evolution has been con-
strained, or modulated, by multiple environmental factors, and a single trait may
respond differently across species to the same stimuli because of this. Here we will
favor the view that fire has been a major driver of plant evolution and, in particular,
of the evolution of physiological traits associated with fire survival and post-fire
recruitment as we will develop below. However, we will also argue that the response
has been modulated, at least to some degree, by drought and temperature stress and
other environmental factors (Fig. 4.1). We note that adaptive responses to fire
include direct responses to the heat shock, but also indirect responses from smoke
or combustion products. We will also discuss the possibility, raised by some authors,
on whether or not plant evolution has led to specific adaptations that enhance flam-
mability or that reduce it (i.e., anti-flammability). The definition of flammability
will be discussed in the relevant section of this chapter (Sect. 4.5).
It is nearly impossible to obtain direct evidence on whether a trait has evolved in
response to fire. Consequently, discussions on whether traits favoring post-fire suc-
cess are indeed fire adaptations or not are often based on a certain degree of specula-
tion. They are, for instance, based on the temporal matching between the first
appearance of the adaptation and the degree of fire activity at the time (Fig. 4.2).
That is, if a trait first appeared at a time of high fire activity, it is then considered as
a fire adaptation. In fact, peaks in wildfire activity in the geological record have co-­
occurred with peak speciation events (Fig. 4.2). Similarly, if a trait only appears in
fire-prone environments, it is also considered as fire adaptation. This approach is
problematic because the evidence is only circumstantial.
It is important to remember that plants are adapted to a given fire regime, and not
to fire per se, as discussed in Chap. 2. This means that traits providing a fitness
advantage over a certain fire regime may be maladaptive under a different fire
regime. The expression “plants being adapted to fire” in this book should thus be
understood as a shortcut to indicate that plants have adapted to a certain fire regime
(see Chap. 2).
Here we first present the genetic mechanism that could underlie plant evolution
under fire. That is, if fire is a catalyzer of plant evolution, there must be an ­underlying
4.1 Introduction 55

Fig. 4.1 Distinguishing adaptations from exaptations. Continuous, dotted, and dashed lines indi-
cate when fire, fire along with another environmental pressure, or an environmental pressure dif-
ferent from fire, have shaped the value of the trait. In 1, the value of the trait does not change
temporally, and no adaptation or exaptation consequently occurs. In 2, fire is not present, but there
are other environmental pressures. Consequently, the changes in the trait’s fitness under fire (left
column) represent an exaptation in 2a and no adaptation in 2b. Contrastingly, changes in the right
column represent an adaptation to such prevailing environmental pressure in both 2a and 2b. In 3a
and 3c, fire is present, and in 3b and 3d, fire and other environmental pressures concur. The trait’s
fitness increases under fire (left column, a and c) and under fire + other environmental pressures (b
and d). Consequently, this also represents fire adaptations, but in some cases (b and d), the evolu-
tion of the trait in that direction has not been shaped only by fire but by fire and other environmen-
tal stimuli. In 3c and 3d, we observe trade-offs. That is, the trait’s fitness value increases under fire
(left column), but it diminishes under other evolutionary pressures (right column), indicating a
trade-off. In 4, we recognize that changes do not need to be constant through time and that adapta-
tions may occur (in b, c, and d), even if the pressure from fire, or from fire and other environmental
pressures, only acted for a part of a plant’s evolutionary history. Note that there are other evolution-
ary routes not depicted in this graph

mechanism that drives or catalyzes the process, and mutagenic properties along with
selection may provide such mechanisms. Then, we will discuss the evidence on the
potential role of fire as a driver of the evolution of plant regeneration strategies,
allocation patterns to enhance post-fire survival, and (anti-)flammability traits
(Tables 4.1 and 4.2). We will finally discuss some trade-offs between adaptations to
fire and to other environmental stressors and, particularly, drought.
56 4 The Evolution of Physiological Adaptations in a Flammable Planet

Fig. 4.2 Geological time chart, temporal evolution of fossil charcoal content (an indicator of fire
activity), conifer and angiosperm diversification, and associated major vegetation events. Time
series data on fossil charcoal was digitized from Glasspool and Scott (2010), and data on conifer
and angiosperm speciation rates were digitized from He and Lamont (2018a), and it shows relative
values (actual value divided by the maximum)

4.2 Fire as a Selective and Mutagenic Agent

Wildfires may alter evolution by selecting for (or against) certain phenotypes. There
are some well-documented examples of this process. For instance, seeds of the
shrub Helenium aromaticum showed higher pubescence, width/length ratio, and
pericarp thickness in areas with more frequent fire regimes. After exposing this spe-
cies to an experimental fire, there was a higher germination of hairier, less rounded
seeds and those with a thicker pericarp (Gómez-González et al. 2011). These authors
also demonstrated that variation in seed pubescence and shape (but not in pericarp
thickness) is heritable among progenies, thus showing how fire could be an agent
underlying rapid trait evolution. How general is this rapid evolution in response to
fire remains to be understood, and it is not always found in species from fire-prone
environments (Torres et al. 2018).
Fire may also alter trait evolution by its mutagenic effects. The mechanisms
explaining the role of fire as a mutagenic agent were developed by He and Lamont
(2018b), who argued that fire-induced evolution is driven by heat and chemical
4.2 Fire as a Selective and Mutagenic Agent 57

Table 4.1 Traits related to persistence or recruitment whose evolution is likely to have been
shaped by fire
Possible
evolutionary
Function Examples Likely evolutionary pressure pressure
Enhance Post-fire seed Fire, drought, heat Nutrients,
recruitment release (serotiny) herbivory
Post-fire seed Fire, drought, heat, herbivory Nutrients
germination
Post-fire Fire Other
flowering disturbances
Enhance Resprouting Any agent consuming biomass (but Any agent
persistence resprouting from epicormic buds and deleting biomass
lignotuber may be primarily in response
to fire)
Meristem Fire may result in thickening of outer Inner bark linked
protection bark to metabolism
We also note other environmental pressures, apart from fire, that are either likely or possible to act
upon trait evolution

Table 4.2 Traits potentially increasing or decreasing flammability from leaf to plant scale
Leaf scale Plant scale
Trait type Example Trait type Example
Leaf Leaf carbon, nitrogen, and Light interception Leaf angle, branching
economics phosphorous, specific leaf area patterns, bulk density
Defenses Resins, volatiles, waxes, ash Conductance and Leaf size and shape
content energy balance
Drought Moisture content Light/shade Height to first branching,
tolerance retention of dead matter
Phenology Leaf flushing and
shedding, photoperiod
sensitivity

effects. Heat may act as a mutagenic agent by altering meiosis and inducing diploid
gametes which, in turn, may lead to polyploidy and genome duplication. Polyploidy
is one of the main mechanisms leading to speciation in plants, and it increases the
environmental range a plant may tolerate (Adams and Wendel 2005; Vanneste et al.
2014). There is some circumstantial evidence supporting this hypothesis, since
polyploidy has been documented as a mechanism that allowed plant survival during
the Cretaceous-Paleocene extinction event. This is a time with high fire activity and
also the time when genome duplication in Myrtaceae and Proteaceae, two Australian
and South African families of fire-adapted trees, is expected to have occurred (He
et al. 2019).
However, it still needs to be demonstrated whether fire-released heat does affect
meiosis in such a way. That is, experiments demonstrating that heat during meiosis
increases gamete duplication are still rare and they have been conducted after
58 4 The Evolution of Physiological Adaptations in a Flammable Planet

a­ pplication of a mild and continuous heat (~36 °C) for a prolonged period of time
(1–2 days) (Pecrix et al. 2011). The thermal profile that occurs during a fire is very
different from such pattern, and it remains to be demonstrated under which condi-
tions can fire-released heat affect meiosis.
Fire may also act as a mutagenic agent due to the chemical elements present in
smoke. In particular, polycyclic aromatic hydrocarbons (PAHs), a well-known com-
pound of smoke, may be absorbed by the surviving plant parts and consequently
lead to mutations, and it may also affect directly the mature seed. Interestingly, He
et al. (2019) state that PAH extracts in species from fire-prone environments are
more mutagenic than those from species where fire is less common. However, this
statement followed from a comparison across different studies and potential effects
arising from different methodologies used in the different data sources was not
considered.
Combustion by-products include different solids that are deposited on the soil,
and they can be potentially either taken up by the roots or penetrate the seed or
seedling. These by-products include different substances with mutagenic properties
such as metal oxides and radioisotopes as well as synthesized organic compounds,
constituting the third mechanism by which fire could act as a mutagenic agent (He
and Lamont 2018b).

4.3  ire and the Evolution of Strategies Enhancing


F
Recruitment

4.3.1 Serotiny

The earliest plant adaptation to fire that has been described so far is serotiny, the
accumulation of seeds in the canopy that are released after some environmental
stimuli. While the anatomical features may vary across species, serotiny often
requires that the resins’ sealing follicles or scales melt with the heat, which subse-
quently allows for seed release. According to He et al. (2016), serotiny appeared for
the first time in the Carboniferous, 332 mya, in early conifers. It is difficult to infer
whether a cone was serotinous from the fossil record, and the authors made a num-
ber of assumptions for deciding whether a species was or not serotinous. Their
effort is laudable, but independent validation of their definition of serotiny will be
required to confirm these findings. Other undisputed records of serotiny have been
reported for the Cretaceous for some extinct Cupressaceae (Mays et al. 2017) and
Pinus (He et al. 2012), and in the Paleocene for Banksia (He et al. 2011), with more
recent families also showing serotiny.
Serotiny is common in Mediterranean-type ecosystems (MTEs) from Australia
and South Africa, and it is also found, but rarely, in Northern Hemisphere conifers.
While serotiny might have originally evolved in response to fire, that does not pre-
clude that other environmental factors such as drought (Axelrod 1980) or nutrient
4.3 Fire and the Evolution of Strategies Enhancing Recruitment 59

scarcity (Bradshaw et al. 2011) may have also played a role. According to these
authors, nutrient scarcity could favor the evolution of serotiny when resources are
so scarce that only a few seeds are produced. Indeed, Australian soils are putatively
P limited and MTE soils are, overall, poor. Under such circumstances, the cohort of
seeds produced within a year could be small, and herbivores could eat the few seeds
produced within a year. Consequently, it would be advantageous to accumulate the
seeds produced over a few years in the form of aerial seed banks. However, seed
masting, a widespread phenomenon where seed production is synchronous and
cyclic within populations (Herrera et al. 1998), would suffice as an adaptation under
these conditions (Kelly and Sork 2002). The argument that serotiny has evolved
under nutrient scarcity is thus speculative, and further testing is still necessary.
There is more evidence that the evolution of serotiny has been shaped by drought.
Indeed, it is common for many species to release seeds after drought or excessive
exposure to sunlight. P. halepensis is a species that originated in the drier Miocene
(Gallien et al. 2016). It is common in the Western Mediterranean Basin, and it shows
xeriscence (seed release after drought) in addition to pyriscence (seed release after
fire). Serotiny in this species thus seems to follow a bet hedging strategy, rather than
a strict fire adaptation. That is, only some cones open in response to a certain stimuli
(drought or fire), while others stay closed and will open later, after subsequent stim-
uli. Serotiny in conifers may have first evolved in response to fire. However, in the
case of P. halepensis, a species that first appeared in a time of drought, it is conceiv-
able that such drought and other factors later modified serotiny to achieve both
xeriscence and pyriscence as an adaptation to multiple concomitant stresses such as
fire or drought (line 4a in Fig. 4.1).
Support from this hypothesis comes from various studies indicating that the degree
of serotiny (the proportion of serotinous cones relative to the total number of cones)
is lower in provenances originating from dry environments (Hernández-­Serrano et al.
2014). That is, studies that have examined genetic variation in the degree of serotiny
in a P. halepensis genetic trial, where individuals from different geographic origins
were growing together, have observed that populations from drier environments show
a higher degree of serotiny, a largest investment into reproduction, and they also
started to reproduce earlier than progenies from wetter environments (Fig. 4.3).
Some authors have interpreted this result as evidence for serotiny as a fire adapta-
tion, by assuming that fire activity increases linearly with aridity (Hernández-­
Serrano et al. 2014). However, it is more plausible to interpret this pattern as a direct
response of serotiny to aridity. Furthermore, the assumption that fire activity
increases linearly with drought in Mediterranean environments contradicts observed
patterns of fire activity. Fire activity in the Mediterranean follows a unimodal pat-
tern, peaking at intermediate precipitation levels and declining as aridity increases
(Fig. 4.3). Consequently, current evidence indicates that the higher degree of sero-
tiny with increasing aridity, when water scarcity is highest and fire activity lowest,
is not only an adaptation to fire but to increased aridity.
Lamont et al. (2019) argue that serotiny is a strong fire-selected trait because,
even if seed release may occur in the absence of fire, post-fire population ­recruitment
is much larger than in inter-fire intervals. However, this argument is compatible with
60 4 The Evolution of Physiological Adaptations in a Flammable Planet

Fig. 4.3 Serotiny is negatively correlated with aridity, measured as summer precipitation (a). In
turn, the relationship between fire activity (indicated by the burned area fraction) and the aridity
index is unimodal, and fire activity declines towards the drier end (b). Gray points in (b) indicate
the mean annual burned area fraction (1996–2012) and the aridity index (the ratio between mean
annual potential evapotranspiration to precipitation). European points are highlighted in blue (grid
cells, 0.25°× 0.25°), and the line indicates the 0.99 quantile. (Data in the left panel from Hernández-­
Serrano et al. (2014). Right panel redrawn with permission from Karavani et al. (2018))

the notion that serotiny evolved in response to fire and that it was later modulated by
drought or other environmental factors. Furthermore, because there are different
degrees of serotiny across, and also within, species, the relative importances of
inter- vs post-fire recruitment are likely to vary. It is important to remember that
serotiny is considered as an adaptation to crown fire regimes, but some serotinous
species also develop thick barks (Fernandes et al. 2008), which are indicative of
surface fire regimes. Xeriscence could thus enhance recruitment after a surface fire.
We argue that further progress in the area will arise from fully embracing the com-
plexity occurring in ecosystems: different fire types co-occur, and plants need to
survive under different kinds of stresses and perturbations, such that bet hedging
strategies that are under multiple evolutionary pressures are more likely to emerge.

4.3.2 Seed Germination

Another regeneration trait that could have evolved in response to fire activity is seed
germination in those species that accumulate seeds in soils and develop soil seed
banks. In such species, seeds enter dormancy, and they require some physical or
chemical stimuli to recover physiological activity. Some extreme examples include
Emmenanthe penduliflora, which only germinates in the presence of combustion
products (Wicklow 1977). Fire could break seed dormancy when the heat cracks the
cuticle or outer tissues such that imbibition is allowed. Fire could also break seed
dormancy chemically with the compounds occurring in the smoke, as some may
stimulate seed germination. As with serotiny, current evidence indicates that seed
germination is more likely to have evolved in response to fire and other stresses such
as drought or herbivory, rather than in response to fire alone.
4.3 Fire and the Evolution of Strategies Enhancing Recruitment 61

A hard seed coat that requires thermal scarification may also crack after passage
by a digestive tract. It is thus difficult to understand whether germination induced by
heat shocks is an adaptation to fire or to herbivory. In order to separate them, some
authors have compared the responses across plants from contrasting geographical
origins, under the expectation that germination would be enhanced after heat shock
application only in species from fire-prone environments. For instance, Luna et al.
(2007) compared the responses across 57 species that were from the Iberian
Peninsula, from the Mediterranean Basin, or with wide distribution ranges. They
observed how these species were tolerant to the heat shock. However, they did not
observe that thermal scarification promoted germination, and on the contrary, ger-
mination declined under high temperatures for all geographic origins. These authors
concluded that seed germination may occur under conditions that lead to soil tem-
peratures of up 80 °C, such as briefly after a light surface fire or by high summer
temperatures. However, germination declined when temperatures reached over
100 °C, which would be those occurring during a high-intensity fire.
There are some patterns consistent with the notion that a hard coat requiring
thermal scarification is a fire adaptation, such as the fact that germination in resprout-
ers has sometimes been reported as being more sensitive to high temperatures than
germination in seeders (Luna et al. 2007). However, this response is also confounded
by life-form, as some functional types are more sensitive to the heat shock than oth-
ers (Ne’eman et al. 2012). Therefore, the hypothesis that thermal scarification as a
requirement for germination is exclusively a fire adaptation does not explain these
patterns. It is more plausible to assume that hard coats are a bet hedging strategy
that evolved under fire, herbivory, and high soil temperatures.
Some authors suggest that smoke-stimulated seed germination originated during
the Cretaceous in the Proteaceae, Haemodoraceae, and Restionaceae. He and
Lamont (2018a) reviewed the evolution of smoke as a mechanism signaling for seed
germination, and they proposed that this was the ancestral condition for seed flower-
ing plants. Loss of smoke sensitivity, or the development of a hard coat, would
originate later.
Among the different elements in smoke, karrikins are considered as the com-
pounds that stimulate germination. In fact, it has been often documented that kar-
rikins act as general stimulants for germination, and they have also been observed
in plants that do not occur in fire-prone environments. Bradshaw et al. (2011) argued
that this is because the stimulant for seed germination is present in soils as a by-­
product of microbial oxidation of organic matter, and consequently, smoke-induced
germination is not a fire adaptation. However, He and Lamont (2018a) explain the
apparent universality of karrikin-induced germination given its ancestral condition
in flowering plants. That is, they consider that plants from no-fire environments with
karrikin-induced germination have developed from ancestors that occurred in fire
environments.
Wildfires and drought often coexist, and while water scarcity reduces seed ger-
mination, smoke partly compensates for such a decline. That is, smoke reduces the
sensitivity of germination to water scarcity, such that post-fire germination may
occur under a wider range of drought environments in some species (Thomas et al.
62 4 The Evolution of Physiological Adaptations in a Flammable Planet

2010). However, these results do not seem to hold for shrubs with hard-coated
seeds. Studies on these species have reported that smoke either does not alter the
water stress sensitivity of germination (Chamorro and Moreno 2019) or that smoke
further inhibits germination at high levels of water scarcity in Mediterranean hard-­
coated shrubs (Chamorro et al. 2017). That germination under high water scarcity
was further reduced by smoke cannot be viewed as a fire adaptation. Wildfires co-­
occur with drought in many forests and shrublands and a key to post-fire success is
early germination (de Luis et al. 2008; Karavani et al. 2018). Any factor that delays
germination, such as smoke under drought, would thus be maladaptive (Chamorro
et al. 2017). However, the generality of such interactive effect still needs to be
demonstrated.
There is also evidence for maternal effects as plants subjected to drought produce
seeds with smaller germination rates, and smoke cues further reduce final germina-
tion (Chamorro et al. 2016). At any rate, the jury is thus still out as to how fire and
drought have shaped seed germination.

4.4  ire and the Evolution of Strategies Enhancing


F
Persistence

4.4.1 Resprouting

Resprouting has traditionally been considered as the ancestral trait in angiosperms


(Wells 1969; Bond and Midgley 2003), while seeding was viewed as a posterior
adaptation, although recent research has challenged this concept (Lamont et al.
2019). However, being able to resprout does not imply that a species may resprout
after fire. Unlike other disturbances, fire consists of a removal of biomass along with
thermal stress near the tissues directly consumed. The ability of post-fire resprout-
ing is thus linked with the location of the buds, so they must show some degree of
thermal protection to survive after the fire (see Chap. 7). Furthermore, not all types
of resprouting have been linked exclusively to fire activity. For instance, resprouting
from roots, rhizomes, or the root collar may be a response to other disturbances such
as herbivory. However, some authors consider that resprouting from epicormic buds
or lignotubers is linked to fire (Pausas and Keeley 2014).
Lamont et al. (2019) reconstructed the origins of post-fire resprouting for differ-
ent clades. They documented that all of the examined clades originated from non-­
fire-­prone ancestors, with the implication being that resprouting as an ancestral trait
would not have been fire-induced. Post-fire resprouting became increasingly com-
mon since the Cretaceous (Fig. 4.4). Lamont et al. (2019) also compared the appear-
ance of new resprouting lineages in areas currently experiencing Mediterranean
climates, with fires often occurring in summer and fall, with savanna grasslands,
where natural fires are mostly confined to the winter and spring (Fig. 4.4). Although
Mediterranean climate is relatively recent, probably appearing during the Miocene,
fire-induced resprouting appeared at least 100 mya in this area. There was an
4.4 Fire and the Evolution of Strategies Enhancing Persistence 63

Fig. 4.4 Appearance of


post-­fire-­resprouting
lineages, or of novel
resprouting types, over the
last 100 mya for clades
with data available (from
Lamont et al. 2019). Each
line represents either areas
currently experiencing fires
in summer-autumn (mostly
Mediterranean climates) or
with fires in winter-­spring
(like savanna grasslands)

increase in the number of resprouting lineages ~55 mya, roughly coinciding with
the 2–4 °C warming at the Paleocene-Eocene boundary. Another increase in the
appearance of new fire-resprouting lineages started at the onset of the Miocene, and
new resprouting lineages have continued to appear until nowadays. In contrast,
areas currently occupied by savanna grasslands only started to show resprouting
lineages in the last 20 mya (Fig. 4.4). Here, the initial start of the appearance of
resprouting is considered as a period with low fire activity, and it precedes the onset
of the savanna biome, and the consequent increase in burned area, that started
~8 mya in the area. Fires in the savanna biome show very short return intervals, typi-
cally appearing every less than 5 years and sometimes every year. Such short return
interval fires favor resprouters over seeders. In the Mediterranean, however, fire
occurs at decadal scales. It is thus to be expected larger evolutionary pressure for
appearance of resprouters over the savanna biome than in the Mediterranean.
We can only speculate under which conditions is resprouting favored over seed-
ing. Current hypotheses propose that interactions between the fire regime and pro-
ductivity favor one regeneration mode over the other. Post-fire seeding requires
stand-replacing fires that kill the original stand. It requires intermediate fire inter-
vals. That is, intervals where there is enough time for individuals to reach maturity
and produce seeds, but also where the fire interval is shorter than the individual
lifespan. This is particularly the case for species that keep their seeds in their can-
opy, as all seeds would be lost if the fire occurs after the mother plants have per-
ished. Resprouters will often be able to survive under extremely short, long, or
erratic fire intervals, since obligate seeders may not be able to recruit under these
circumstances. However, intermediate fire intervals allow for coexistence of both
resprouters and seeders.
Under these intermediate fire intervals, some authors consider resprouters to be
favored at productive sites. This is because resprouting is considered to be a more
64 4 The Evolution of Physiological Adaptations in a Flammable Planet

expensive strategy: it requires storing carbohydrates and developing and protecting


a bud bank (Pausas and Keeley 2014). Seeders could be more frugal species, but
they show longer post-fire establishment times: they need to germinate and develop
a new rooting system after the fire, while resprouters already have a well-developed
root system, so they are “ready to go.” However, the biogeography of resprouters is
still disputed, and other authors consider resprouting to be favored under less pro-
ductive environments (Cruz and Moreno 2001; Ojeda 1998; Midgley 1996).
In fact, differences in these hypothesized costs associated with seeding and
resprouting imply that seeders are favored under higher aridity and resprouters
under more productive environments (Pausas and Keeley 2014). Seeding may be
further complicated in sites where seed regeneration is more difficult, such as rocky
terrains (Lamont et al. 2019). However, support for these hypotheses is mixed. After
crossing different datasets, we can observe that seeders tend to occur over shallower
soils than resprouters, but differences are not significant (Fig. 4.5). There are also no
differences over the mean annual precipitation (the main driver of productivity
globally) of sites where both functional types tend to occur (Fig. 4.5). It is also
unclear whether increasing carbon allocation to storage or buds represents a sub-
stantial disadvantage for resprouters, because carbohydrate concentrations seldom
limit plant growth (Korner 2003). Thus, the hypothesis that higher resprouting costs
imply that these species dominate more productive sites does not seem to be sup-
ported by current evidence.
Seeders and resprouters show very marked functional differences to survive
under drought, with seeders showing a tolerant strategy (shallow roots and xylem
resistant to embolism) and resprouters seeking to avoid drought stress (with deep
tap roots that allow for larger soil water exploration) (Vilagrosa et al. 2014). As
noted by Parra and Moreno (2017), post-fire environments are less exposed to water
stress: evapotranspiration declines and competition for water is negligible. We will
examine more fully the functional differences across both regeneration types in sub-
sequent chapters (Chaps. 6 and 7). Of relevance here is that there is little evidence
that one regeneration mode is better adapted than the other to survive under drought
or across productivity gradients (Fig. 4.5).
Seeders and resprouters do tend to occupy different successional niches, at least
in Mediterranean ecosystems. Resprouters tend towards showing higher shade tol-
erances (Chap. 6, Fig. 6.2), whereas seeders often show higher light requirements.
Under intermediate fire frequencies, it is thus likely that both regeneration modes
are complementary and that they have simply evolved to occupy niches that would
otherwise remain empty.

4.4.2 Meristem Protection

Thin barks are often considered as the ancestral trait, and bark thickening is consid-
ered as a posterior adaptation. Barks provide thermal insulation, and as rule of
thumb, barks thicker than 1 cm protect from light surface fires and barks thicker
4.4 Fire and the Evolution of Strategies Enhancing Persistence 65

Fig. 4.5 Differences on


environments occupied,
and drought tolerances
shown, by woody
resprouters (R+) and
seeders (R−). Data were
analyzed statistically with
an ANOVA, and significant
differences (at P < 0.05)
between R− and R+ did
not occur for any trait.
Data on whether woody
species resprout or not
after fire comes from
Tavşanoğlu and Pausas
(2018), on soil depth and
precipitation from Maire
et al. (2015), and on
drought tolerances from
Niinemets and Valladares
(2006)

than 2 cm protect cambium from higher-intensity surface fires (see Chap. 7 for the
relationship between bark thickness and survival). Bark thickening for bud and mer-
istem insolation is considered one of the earliest adaptations to fire. In early coni-
fers, 15-mm-thick barks have been documented for Protopitys buchiana, at an
estimated age of 359–347 mya (Decombeix 2013). In Pinus, barks thicker than
15 mm have been estimated to appear 105–147 mya, 40 million years earlier than
serotiny. This is considered to be a time with low-intensity surface fires. Barks
thicker than 30 mm have been dated from 80 to 96 mya (He et al. 2012), at a time
where more severe surface fires may have occurred (Bond and Midgley 2012). Crisp
et al. (2011) documented that major changes in bark thickening, and related bud
location, occurred in Myrtaceae starting at the Paleogene, 62 mya.
Variations in bark thickness have often been examined from the lens of fire pro-
tection. However, bark serves a series of additional functions, including photosyn-
thate transport; stem photosynthesis; reserves storage; protection against pathogens,
herbivores, and high temperatures; wound closure; and mechanical support, among
66 4 The Evolution of Physiological Adaptations in a Flammable Planet

others (Rosell 2019). It is thus likely that bark has been subject to multiple evolu-
tionary pressures and that its evolution has not only been shaped by fire.
The bark is divided between the inner and the outer sections. The inner bark is
composed of living cells, and it is thus responsible for the metabolic functions pre-
viously mentioned. The outer bark is composed of dead tissue, and its main function
is to protect the living inner bark, cambium and xylem. In order to understand to
what extent is bark thickness driven by fire activity, Rosell (2016) sampled 640 spe-
cies across 18 communities representative of the major biomes globally, and across
highly contrasting climatic and fire regimes. She observed that the major driver of
bark thickness across plant species is stem diameter (r = 0.72), and the relation
between inner bark thickness and stem diameter was closer (r = 0.65) than between
outer bark thickness and stem diameter (r = 0.35). The strong correlation between
stem diameter and bark thickness, and the fact this correlation is mostly driven by
the inner bark thickness, is consistent with the hypothesis that metabolic support is
the main function underlying variation in bark thickness.
Because the inner bark serves metabolic functions, the effects of fire activity are
more likely to drive thickening of the outer bark, but even then it is difficult to sepa-
rate its effects from other damages such as herbivores. Furthermore, Rosell (2016)
observed limited evidence that the ratio of inner to outer ratio is affected by fire
activity as the relationship of this ratio and fire return interval is not significant
(Fig. 4.6) across communities. There is no doubt that thicker barks protect from fire.
However, understanding to what extent has fire driven bark thickening remains
unresolved.

4.5 Fire and the Evolution of (Anti-)flammability

Perhaps the most controversial aspect on how fire has shaped plant evolution is plant
flammability. Current discussions on landscape flammability are rooted on two
seminal papers that appeared in 1970. On the one hand, Anderson (1970) proposed
a definition of leaf flammability based on the time to ignition (ignitability), how
well the fire burns endogenously (sustainability), and how fast a fire burns (combus-
tibility). This definition, along with a modification introduced by Martin et al.
(1994) to acknowledge consumability, forms the basis of today’s definition of leaf-­
scale flammability.
On the other hand, Mutch (1970) argued that plant communities from fire-prone
environments have evolved to become more flammable than those from fire-­
independent communities. This hypothesis is now discredited because it assumed
group selection. However, it served as inspiration for later hypotheses proposing
individualistic arguments on the evolution of flammability. The “Kill thy neighbor”
hypothesis (Bond and Midgley 1995) proposes that traits enhancing flammability
could evolve in species with seed banks that regenerate after fire: competition from
other species would diminish in post-fire environments, and this would favor the
recruitment of its own offspring. Gagnon et al. (2010) subsequently proposed the
4.5 Fire and the Evolution of (Anti-)flammability 67

1.0
0.8
Inner/outer bark ratio
0.6

Arid shrubland
Cool sclerophyll forest
Cool temperate woodland
0.4

Desert
Mediterranean shrubland
Mediterranean woodland
Savanna
0.2

Seasonally dry forest


Temperate deciduous
Temperate rainforest
Temperate sclerophyll forest
Temperate woodland
0.0

Tropical rainforest

0 20 40 60 80 100
Fire return interval (years)

Fig. 4.6 Ratio of inner to outer bark ratio as a function of fire return interval for 640 species across
18 communities. Values indicate median and error bars the first and third quantiles. (Data from
(Rosell (2016))

“pyrogenicity as protection” hypothesis that included post-fire seeders and also


resprouters. It argues that there could have been selection towards traits enhancing
crown flammability and that protect belowground organs from fire. More recently,
Pausas et al. (2017) proposed that whole-plant flammability depended on the inter-
actions between combining ignitability, combustibility, and consumability. These
authors considered that flammability could have evolved in order to lead to different
fire types (e.g., surface fires, grass fires, or crown fires) depending upon the post-fire
regeneration strategy.
This so-called adaptationist program has been met with criticism by some scien-
tists. Bradshaw et al. (2011) considered that definitions of adaptations needed to
more rigor, so as not confuse them with exaptations. Midgley (2013) criticized the
“Kill thy neighbor” and the “pyrogenicity as protection” hypotheses on the basis of
the underlying assumptions regarding patterns of seed dispersal and fitness advan-
tages. He concluded that flammability is a property that may be incidental or that
may emerge in species and ecosystems, but that evolution towards higher flamma-
bility could not occur. On the contrary, he hypothesized that only evolution towards
lower flammability could arise. Bowman et al. (2014) similarly concluded that it is
extremely difficult to demonstrate that species have evolved to “self-immolate” and
that it is more plausible to assume that plants have evolved mechanisms to survive,
rather than to start, fires.
68 4 The Evolution of Physiological Adaptations in a Flammable Planet

Beyond the work from these authors, there are major criticisms to the hypothesis
that plants have evolved traits enhancing flammability. The first is the definition of
flammability beyond the leaf scale. We are still lacking a clear definition of flam-
mability because it is not exclusively a fuel property: it varies with environmental
conditions, and to which degree do physiological or functional traits affect fire
behavior is unclear. We will more fully discuss to what degree do fuel attributes
drive fire behavior in Chap. 9.
The second problem with the “adaptationist program” is that there are so many
plant traits potentially affecting fire behavior (Table 4.2), that one must then accept
that fire has been a major driver of the evolution of plant traits associated with the
leaf economics spectrum, defenses, drought tolerance, light interception, conduc-
tance and energy balance, light and shade tolerance, and phenology (Table 4.2). We
discuss in Chap. 6 the physiological function of these traits in relation to wildfires.
Of relevance here is that, while some fuel attributes may influence the capacity of a
plant to burn, accepting that all traits potentially affecting fire behavior have been
selected by fire may be taking things a bit too far. Assuming that flammability
emerges as a result of the changes in the parameters mentioned in Table 4.2 is thus
more plausible (Midgley 2013).
Some aspects of plant structure may hinder fire activity, such as thick and corky
barks, high moisture content, and a high degree of self-pruning. Consequently,
some authors consider that evolution for traits decreasing flammability (anti-­
flammability) may have occurred (Midgley 2013). However, this argument suffers
from the same problems as those outlined previously for the argument that flam-
mability has been selected for.

4.6 Adaptations to Fire and Physiological Trade-Offs

We have argued through this chapter that plant physiological evolution has been
shaped by interactions between fire regimes, environmental stressors, and plant
metabolic demands. All the examples considered so far have assumed that stress and
perturbation drive trait evolution towards the same direction. However, the opposite
may also occur. There are long-standing ecological theories proposing that plants
may survive under stress or perturbation, but not under both (Grime 1977; Pierce
et al. 2017).
This has been, for instance, documented for conifers, where one may encounter
species adapted to high aridity or to high fire activity, but not to both stress and
perturbation concomitantly (Fig. 4.7). Considering how fire activity varies across
productivity gradients (Fig. 4.3), Keeley (2012) proposed a division of life strate-
gies across Pinus species in relation to fire as tolerant, embracer, and avoider spe-
cies. Resco de Dios et al. (2018) later modified this classification to include
resistance to embolism, an indicator of drought tolerance.
Fire-tolerant conifers occur at the more productive sites, where fire activity is
often limited by high moisture that results in low-intensity surface fires (Fig. 4.7),
4.6 Adaptations to Fire and Physiological Trade-Offs 69

4.0
+ ● ●● ●

Tolerant P < 1e−04

Fire Tolerance
Site productivity

3.0
● ● ● ●

Embracer

2.0
● ● ●●●●●●● ● ●● ●●●● ● ●

Avoider
-

1.0
●●●
●● ●● ● ●

+ Fire tolerance - 0 −4 −8 −12


Y50 (MPa)
- Embolism resistance +

Fig. 4.7 Trade-offs between resistance to fire and drought. A simplified view on the distribution
of conifer species across productivity gradients may be hypothesize to be shaped by fire and
drought (left). Evidence for the trade-off comes from analyzing, for instance, the degree of fire
tolerance (PLANTS USGS Database) and the pressure potential at which 50% of hydraulic con-
ductance is lost (Ψ50, (Choat et al. 2012)), an indicator of embolism resistance and, consequently,
of drought tolerance. The line in the left panel indicates the results of quantile regression. (Redrawn
with permission from Resco de Dios et al. (2018) and Karavani et al. (2018))

and species do not show high embolism resistance. Consequently, fire-tolerant spe-
cies (e.g., P. nigra, P. ponderosa, P. sylvestris) have thick barks that allow the sur-
vival of individuals under low-intensity fires, do not regenerate under high-intensity
crown defoliating fires, and are sensitive to drought. Fire-embracer species occur at
sites with intermediate productivity, where fire activity is highest, and have a low
degree of self-pruning and an overall canopy architecture that enhances crown fires
(Fig. 4.7). Their regeneration depends on stand-replacing fires that open their seroti-
nous cones (e.g., P. attenuata, P. halepensis), and they show intermediate resistance
to embolism. Fire-avoider species occur at dry (or also upper montane) environ-
ments, where fires are very rare and limited by fuel load (or moisture), lack adapta-
tions to fire (e.g., P. aristata, P. uncinata; Fig. 4.7), and show the highest tolerance
to drought.
Evidence for other trade-offs exist, including one between resistance and recruit-
ment after fire. For instance, drier populations of P. halepensis show a relatively
lower investment to bark and a relatively higher investment to reproduction, relative
to those from wetter provenances (Martin-Sanz et al. 2019). The hypothesized
mechanism underlying all these trade-offs could be related to C allocation (Resco
de Dios et al. 2018). Photosynthates may be allocated through the phloem, towards
the xylem, or for increased reproduction. This view, however, is still preliminary,
and further data are necessary to confirm it. Further progress on how fires have
shaped evolution will likely come by studies that examine evolution under multiple
stresses and disturbances. We must show a positive distrust for simplicity (Grubb
1992). Even if it makes conceptual sense that either fire has not shaped plant evolu-
tion at all or that fire is one of main drivers of plant evolution, it is more plausible to
assume that plants need to concomitantly respond to multiple environmental stim-
uli, and consequently, plant traits have evolved in response to such multiple stimuli.
70 4 The Evolution of Physiological Adaptations in a Flammable Planet

References

Adams KL, Wendel JF (2005) Polyploidy and genome evolution in plants. Curr Opin Plant Biol
8(2):135–141. https://doi.org/10.1016/j.pbi.2005.01.001
Anderson HE (1970) Forest fuel ignitibility. Fire Technol 6:312–319. https://doi.org/10.1007/
BF02588932
Axelrod DI (1980) History of the maritime closed-cone pines, Alta and Baja California. Univ Calif
Publ Geol Sci 120:1–143
Bond WJ, Midgley JJ (1995) Kill thy neighbour: an individualistic argument for the evolution of
flammability. Oikos 73:79–85
Bond WJ, Midgley JJ (2003) The evolutionary ecology of sprouting in woody plants. Int J Plant
Sci 164:S103–S114
Bond WJ, Midgley JJ (2012) Fire and the angiosperm revolutions. Int J Plant Sci 173:569–583.
https://doi.org/10.1086/665819
Bowman DMJS, French BJ, Prior LD (2014) Have plants evolved to self-immolate? Front Plant
Sci 5:590. https://doi.org/10.3389/fpls.2014.00590
Bradshaw SD, Dixon KW, Hopper SD, Lambers H, Turner SR (2011) Little evidence for fire-­
adapted plant traits in Mediterranean climate regions. Trends Plant Sci 16(2):69–76
Chamorro D, Luna B, Ourcival JM, Kavgaci A, Sirca C, Mouillot F, Arianoutsou M, Moreno JM
(2017) Germination sensitivity to water stress in four shrubby species across the Mediterranean
Basin. Plant Biol 19:23–31. https://doi.org/10.1111/plb.12450
Chamorro D, Moreno JM (2019) Effects of water stress and smoke on germination of Mediterranean
shrubs with hard or soft coat seeds. Plant Ecol 220(4–5):511–521. https://doi.org/10.1007/
s11258-019-00931-2
Chamorro D, Parra A, Moreno JM (2016) Reproductive output, seed anatomy and germination
under water stress in the seeder Cistus ladanifer subjected to experimental drought. Environ
Exp Bot 123:59–67. https://doi.org/10.1016/j.envexpbot.2015.11.002
Choat B, Jansen S, Brodribb TJ, Cochard H, Delzon S, Bhaskar R, Bucci SJ, Feild TS, Gleason
SM, Hacke UG, Jacobsen AL, Lens F, Maherali H, Martinez-Vilalta J, Mayr S, Mencuccini M,
Mitchell PJ, Nardini A, Pittermann J, Pratt RB, Sperry JS, Westoby M, Wright IJ, Zanne AE
(2012) Global convergence in the vulnerability of forests to drought. Nature 491(7426):752–
755. https://doi.org/10.1038/nature11688
Crisp MD, Burrows GE, Cook LG, Thornhill AH, Bowman DM (2011) Flammable biomes domi-
nated by eucalypts originated at the Cretaceous-Palaeogene boundary. Nat Commun 2:193.
https://doi.org/10.1038/ncomms1191
Cruz A, Moreno JM (2001) Lignotuber size of Erica australis and its relationship with soil
resources. J Veg Sci 12:373–384
de Luis M, Verdu M, Raventós J (2008) Early to rise makes a plant healthy, wealthy, and wise.
Ecology 89:3061–3071
Decombeix A-L (2013) Bark anatomy of an early Carboniferous tree from Australia. IAWA J
34(2):183–196. https://doi.org/10.1163/22941932-00000016
Fernandes PM, Vega JA, Jiménez E, Rigolot E (2008) Fire resistance of European pines. For Ecol
Manag 256(3):246–255. https://doi.org/10.1016/j.foreco.2008.04.032
Gagnon PR, Passmore HA, Platt WJ, Myers JA, Paine CET, Harms KE (2010) Does pyrogenicity
protect burning plants? Ecology 91(12):3481–3486. https://doi.org/10.1890/10-0291.1
Gallien L, Saladin B, Boucher FC, Richardson DM, Zimmermann NE (2016) Does the legacy of
historical biogeography shape current invasiveness in pines? New Phytol 209(3):1096–1105.
https://doi.org/10.1111/nph.13700
Glasspool IJ, Scott AC (2010) Phanerozoic concentrations of atmospheric oxygen reconstructed
from sedimentary charcoal. Nat Geosci 3:627–630. https://doi.org/10.1038/ngeo923
Gómez-González S, Torres-Díaz C, Bustos-Schindler C, Gianoli E (2011) Anthropogenic fire
drives the evolution of seed traits. Proc Natl Acad Sci 108(46):18743–18747. https://doi.
org/10.1073/pnas.1108863108
References 71

Grime JP (1977) Evidence for the existence of three primary strategies in plants and its relevance
to ecological and evolutionary theory. Am Nat 111:1169–1194
Grubb PJ (1992) A positive distrust in simplicity – lessons from plant defences from competition
among plants and among animal. J Ecol 80:585–610
He T, Belcher CM, Lamont BB, Lim SL, McGlone M (2016) A 350-million-year legacy of fire
adaptation among conifers. J Ecol 104(2):352–363. https://doi.org/10.1111/1365-2745.12513
He T, Lamont BB (2018a) Baptism by fire: the pivotal role of ancient conflagrations in evolution of
the Earth’s flora. Natl Sci Rev 5(2):237–254. https://doi.org/10.1093/nsr/nwx041
He T, Lamont BB (2018b) Fire as a potent mutagenic agent among plants. Crit Rev Plant Sci
37(1):1–14. https://doi.org/10.1080/07352689.2018.1453981
He T, Lamont BB, Downes KS (2011) Banksia born to burn. New Phytol 191(1):184–196. https://
doi.org/10.1111/j.1469-8137.2011.03663.x
He T, Lamont BB, Pausas JG (2019) Fire as a key driver of Earth’s biodiversity. Biol Rev Camb
Philos Soc 94:1983–2010. https://doi.org/10.1111/brv.12544
He T, Pausas JG, Belcher CM, Schwilk DW, Lamont BB (2012) Fire-adapted traits
of Pinus arose in the fiery Cretaceous. New Phytol 194(3):751–759. https://doi.
org/10.1111/j.1469-8137.2012.04079.x
Hernández-Serrano A, Verdú M, Santos-Del-Blanco L, Climent J, González-Martínez SC, Pausas
JG (2014) Heritability and quantitative genetic divergence of serotiny, a fire-persistence plant
trait. Ann Bot 114(3):571–577. https://doi.org/10.1093/aob/mcu142
Herrera CM, Jordano P, Guitián J, Traveset A (1998) Annual variability in seed production by
woody plants and the masting concept: reassessment of principles and relationship to pollina-
tion and seed dispersal. Am Nat 152:576–594
Karavani A, Boer MM, Baudena M, Colinas C, Díaz-Sierra R, Pemán J, de Luís M, Enríquez-­
de-­Salamanca Á, Resco de Dios V (2018) Fire-induced deforestation in drought-prone
Mediterranean forests: drivers and unknowns from leaves to communities. Ecol Monogr
88:141–169
Keeley JE (2012) Ecology and evolution of pine life histories. Ann For Sci 69:445–453. https://doi.
org/10.1007/s13595-012-0201-8
Keeley JE, Bond WJ, Bradstock RA, Pausas JG, Rundel PW (2012) Fire in Mediterranean ecosys-
tems- ecology, evolution and management. Cambridge University Press, Cambridge
Kelly D, Sork VL (2002) Mast seeding in perennial plants: why, how, where? Annu Rev Ecol Syst
33:427–447
Korner C (2003) Carbon limitation in trees. J Ecol 91(1):4–17
Lamont BB, He T (2017) Fire-proneness as a prerequisite for the evolution of fire-adapted traits.
Trends Plant Sci 22(4):278–288. https://doi.org/10.1016/j.tplants.2016.11.004
Lamont BB, He T, Yan Z (2019) Evolutionary history of fire-stimulated resprouting, flower-
ing, seed release and germination. Biol Rev Camb Philos Soc 94(3):903–928. https://doi.
org/10.1111/brv.12483
Luna B, Moreno JM, Cruz A, Fernández-González F (2007) Heat-shock and seed germination of a
group of Mediterranean plant species growing in a burned area: an approach based on plant func-
tional types. Environ Exp Bot 60(3):324–333. https://doi.org/10.1016/j.envexpbot.2006.12.014
Maire V, Wright IJ, Prentice IC, Batjes NH, Bhaskar R, van Bodegom PM, Cornwell WK,
Ellsworth D, Niinemets Ü, Ordonez A, Reich PB, Santiago LS (2015) Global effects of soil
and climate on leaf photosynthetic traits and rates. Glob Ecol Biogeogr 24(6):706–717. https://
doi.org/10.1111/geb.12296
Martin RR, Gordon DA, Gutierrez ME, Lee DS, Molina DM, Schoreder RA, Sapsis DB, Stephens
SL, Chambers M (1994) Assessing the flammability of domestic and wildland vegetation. In:
Proceedings of the 12th conference on fire and forest meteorology, SAF Publications, 94-02.
SAF, Bethesda, MD, USA, pp 130–137
Martin-Sanz RC, San-Martin R, Poorter H, Vazquez A, Climent J (2019) How does water avail-
ability affect the allocation to bark in a Mediterranean conifer? Front Plant Sci 10:607. https://
doi.org/10.3389/fpls.2019.00607
72 4 The Evolution of Physiological Adaptations in a Flammable Planet

Mays C, Cantrill DJ, Bevitt JJ (2017) Polar wildfires and conifer serotiny during the Cretaceous
global hothouse. Geology 45(12):1119–1122. https://doi.org/10.1130/g39453.1
Midgley JJ (1996) Why the world’s vegetation is not totally dominated by resprouting plants;
because resprouters are shorter than reseeders. Ecography 19:92–95
Midgley JJ (2013) Flammability is not selected for, it emerges. Aust J Bot 61(2):102–106. https://
doi.org/10.1071/BT12289
Mutch RW (1970) Wildland fires and ecosystems–a hypothesis. Ecology 51:1046–1051
Ne’eman G, Lev-Yadun S, Arianoutsou M (2012) Fire-related traits in Mediterranean basin plants.
Isr J Ecol Evol 58. https://doi.org/10.1560/ijee.58.2-3.177
Niinemets Ü, Valladares F (2006) Tolerance to shade, drought, and waterlogging of tem-
perate northern hemisphere trees and shrubs. Ecol Monogr 76:521–547. https://doi.
org/10.1890/0012-9615(2006)076[0521:TTSDAW]2.0.CO;2
Niklas KJ (2016) Plant evolution – an introduction to the history of life. The University of Chicago
Press, Chicago
Ojeda F (1998) Biogeography of seeder and resprouter Erica species in the Cape Floristic
Region—where are the resprouters? Biol J Lin Soc 63:331–347
Parra A, Moreno JM (2017) Post-fire environments are favourable for plant functioning of seeder
and resprouter Mediterranean shrubs, even under drought. New Phytol. https://doi.org/10.1111/
nph.14454
Pausas JG, Keeley JE (2014) Evolutionary ecology of resprouting and seeding in fire-prone eco-
systems. New Phytol 204(1):55–65. https://doi.org/10.1111/nph.12921
Pausas JG, Keeley JE, Schwilk DW (2017) Flammability as an ecological and evolutionary driver.
J Ecol 105:289–297. https://doi.org/10.1111/1365-2745.12691
Pecrix Y, Rallo G, Folzer H, Cigna M, Gudin S, Le Bris M (2011) Polyploidization mecha-
nisms: temperature environment can induce diploid gamete formation in Rosa sp. J Exp Bot
62(10):3587–3597. https://doi.org/10.1093/jxb/err052
Pierce S, Negreiros D, Cerabolini BEL, Kattge J, Díaz S, Kleyer M, Shipley B, Wright SJ,
Soudzilovskaia NA, Onipchenko VG, van Bodegom PM, Frenette-Dussault C, Weiher E, Pinho
BX, Cornelissen JHC, Grime JP, Thompson K, Hunt R, Wilson PJ, Buffa G, Nyakunga OC,
Reich PB, Caccianiga M, Mangili F, Ceriani RM, Luzzaro A, Brusa G, Siefert A, Barbosa
NPU, Chapin FS, Cornwell WK, Fang J, Fernandes GW, Garnier E, Le Stradic S, Peñuelas
J, Melo FPL, Slaviero A, Tabarelli M, Tampucci D (2017) A global method for calculating
plant CSR ecological strategies applied across biomes world-wide. Funct Ecol 31(2):444–457.
https://doi.org/10.1111/1365-2435.12722
Resco de Dios V, Arteaga C, Hedo J, Gil-Pelegrín E, Voltas J (2018) A trade-off between embolism
resistance and bark thickness in conifers: are drought and fire adaptations antagonistic? Plant
Ecol Divers 11(3):253–258. https://doi.org/10.1080/17550874.2018.1504238
Rosell JA (2016) Bark thickness across the angiosperms: more than just fire. New Phytol
211(1):90–102. https://doi.org/10.1111/nph.13889
Rosell JA (2019) Bark in woody plants: understanding the diversity of a multifunctional structure.
Integr Comp Biol 59(3):535–547. https://doi.org/10.1093/icb/icz057
Tavşanoğlu Ç, Pausas JG (2018) A functional trait database for Mediterranean Basin plants. Scient
Data 5:180135. https://doi.org/10.1038/sdata.2018.135
Thomas PB, Morris EC, Auld TD, Haigh AM (2010) The interaction of temperature, water
availability and fire cues regulates seed germination in a fire-prone landscape. Oecologia
162(2):293–302. https://doi.org/10.1007/s00442-009-1456-0
Torres I, Parra A, Moreno JM, Durka W (2018) No genetic adaptation of the Mediterranean key-
stone shrub Cistus ladanifer in response to experimental fire and extreme drought. PLoS One
13(6):e0199119. https://doi.org/10.1371/journal.pone.0199119
References 73

Vanneste K, Baele G, Maere S, Van de Peer Y (2014) Analysis of 41 plant genomes supports a
wave of successful genome duplications in association with the Cretaceous-Paleogene bound-
ary. Genome Res 24(8):1334–1347. https://doi.org/10.1101/gr.168997.113
Vilagrosa A, Hernandez EI, Luis VC, Cochard H, Pausas JG (2014) Physiological differences
explain the co-existence of different regeneration strategies in Mediterranean ecosystems. New
Phytol 201(4):1277–1288. https://doi.org/10.1111/nph.12584
Wells PV (1969) The relation between mode of reproduction and extent of speciation in woody
genera of the California chaparral. Evolution 23:264–267
Wicklow DT (1977) Germination response in Emmenanthe penduliflora (Hydrophyllaceae).
Ecology 57:201–205
Chapter 5
Environmental Plant Responses
and Wildland Fire Danger

Abstract Fuel moisture is one of the key drivers of fire danger. Fuel moisture limits
combustion during the preheating phase, as water needs to be evaporated before the
fuel can ignite. Some proxies for fuel moisture have been commonly used, like the
Fire Weather Index (FWI) or some of its components, but the success of these prox-
ies has been mixed. It is thus advantageous to incorporate more biological realism
to predicting changes in fuel moisture. Dead fine fuel moisture may be modeled, at
daily scales, from vapor pressure deficit. Live fuel moisture content is driven by
water potential, which is more difficult to forecast at regional scales. It is also pos-
sible to obtain remotely sensed estimates of fuel moisture, but only predictive mod-
els of fuel moisture allow forecasting. Measuring and modeling fuel moisture are
not trivial as different aspects, including variations in live/dead ratio over the fire
season, influence the response. Leaf senescence may indeed lead to large changes in
crown fire likelihood. Although low moisture is a requirement for fire activity,
assuming that lower fuel moisture automatically translates into higher fire danger or
activity may not be correct under some circumstances, as other processes including
atmospheric conditions may play a similar or even larger role.

5.1 Introduction

The moisture content of wildland fuels is one of the components determining wild-
land fire danger and spread (Bradstock 2010; Sullivan 2009). The importance of
fuel moisture is not uniform across the globe, and it is larger in ecosystems with
large fuel loads. Following the classification of fire domains that we will more fully
develop in Chap. 10, fuel moisture plays a dominant role in forests, within the
dryness-­limited fire domain. Ecosystems with large and continuous fuel loads are,
essentially, in a “ready-to-burn” state where fuel moisture will determine its avail-
ability (Boer et al. 2017, 2019). Large fires may occur once fuel moisture falls
below a threshold value in ecosystems with such large fuel loads (Nolan et al.
2016a; Dennison et al. 2008).
Given its importance as a driver of fire danger, some efforts have been developed
to estimate fuel moisture content, and fire danger indices have often been used as
surrogates. Although fire danger indices were not always designed to serve as

© Springer Nature Switzerland AG 2020 75


V. Resco de Dios, Plant-Fire Interactions, Managing Forest Ecosystems 36,
https://doi.org/10.1007/978-3-030-41192-3_5
76 5 Environmental Plant Responses and Wildland Fire Danger

i­ ndicators of fuel moisture, using them to infer fuel moisture has been one of the de
facto applications. This is an unfortunate practice, and the suitability of fire danger
indices as indicators of fuel moisture has often received poor support (Ruffault et al.
2018; Soler Martin et al. 2017; Viegas et al. 2001; Schunk et al. 2017; Wotton and
Beverly 2007; Vinodkumar et al. 2017; Aguado et al. 2007).
Wildland fuels consist of both dead and live plant material. Dead fuels, particu-
larly fine fuels (with a diameter < 25.4 mm), respond rapidly to atmospheric condi-
tions and can be modeled from inputs such as temperature, humidity, or vapor
pressure deficit (Matthews 2013; Viney 1991; Resco de Dios et al. 2015). In con-
trast, live fuel moisture content (LFMC) can be much more difficult to model
because moisture content is a function of plant physiological and structural traits,
which can differ markedly across species, along with variation in soil physical prop-
erties (Nolan et al. 2018; Karavani et al. 2018; Jolly et al. 2014; Nelson 2001;
Macias Fauria et al. 2011).
In order to understand how moisture limits combustion, we begin with a brief
summary of the combustion process. Next, we provide a general description of the
limitations of wildfire danger indices. Then, we will describe the drivers of dead
fuel moisture dynamics and how to obtain spatially explicit estimates. Fourth, we
will discuss the drivers of live fuel moisture and how to develop landscape-scale
estimates of live fuel moisture. Fifth, we will present some problems associated
with measurements of live fuel moisture and how they may bias our current under-
standing. Sixth, we will discuss the evidence on whether fuel moisture is correlated
with fire activity. We will close the chapter with a discussion on whether drought
affects fire activity.

5.2 Fuel Moisture as Limiting Fire Spread

A wildfire results from the combustion of organic matter, and it may be viewed as a
process with four different phases:
Preheating or pre-ignition combustion: this is an endothermic reaction (it requires
energy inputs) where water is evaporated from the fuel. As temperature increases
further, the organic matter begins to decompose releasing volatile gases and
vapors, a process also known as pyrolysis.
Flaming combustion: this is an exothermic and self-sustained reaction where the
flammable gases escaping from the fuel during pyrolysis ignite in the presence of
oxygen, producing light and energy.
Smoldering combustion: this is when the concentration of volatiles is not high
enough to sustain a flame. This usually occurs after the flame has passed and it
often does not contribute to fire spread directly in crown or surface fires. However,
it is the dominant form of fire spread in peat fires.
5.3 Wildfire Danger Indices Provide Limited Information on Environmental Fire Danger 77

Glowing combustion: most volatiles have already burned and there is little smoke,
but the carbon still present in the fuel is being oxidized and heat production
continues.
The fire front advances due to flaming combustion and smoldering and glowing
combustion usually do not contribute to fire spread. Fuel moisture is thus important
in the transition between preheating and flaming combustion, as fuels only ignite
when they are dry enough to burn. Under current oxygen levels, the transition to
flaming combustion is very difficult above the fiber saturation point, at ~30%
humidity (Berry and Roderick 2005), and the ease of ignition grows quickly as
humidity drops below 14% (Watson et al. 1978) (please see Chap. 3 for more
details).

5.3  ildfire Danger Indices Provide Limited Information


W
on Environmental Fire Danger

A variety of wildland fire danger indices have been derived to try to predict the
probability of a fire occurring at a site depending on meteorological conditions
(Keetch and Byram 1968; McArthur 1967; Sharples et al. 2009). Different indices
are used in different parts of the world, but the Fire Weather Index (FWI) is proba-
bly the most popular globally. FWI requires inputs of wind, precipitation, tempera-
ture, and humidity to predict different parameters related to fire occurrence (van
Wagner 1987). The index was originally developed for the boreal forests of Canada,
and it was an index that was ahead of its time in many aspects. In essence, a com-
monality across meteorological weather indices is that they compute a water bal-
ance and they assume that increased fire probability follows after declines in the
water balance. Unfortunately, some indices have been widely implemented before
proper validations were carried out. In fact, different validations on the goodness of
FWI have yielded different results (Fernandes et al. 2016; Urbieta et al. 2015; Bedia
et al. 2015; Abatzoglou and Kolden 2013; Tian et al. 2011; Nogueira et al. 2017;
Flannigan et al. 2016), and FWI did not identify the critical conditions associated
with some of the major fire events in our recent past (Boer et al. 2017).
Regardless of the accuracy of the relation between FWI and fire occurrence,
indices fail to provide a model for seasonal changes in the biological aspects that
impact fire behavior. That is, indices will provide a number, on a scale, that is no
better than an educated guess on fire likelihoods. However, they will not provide any
information on, for instance, the actual fuel moisture of the vegetation. Moreover,
indices are truncated, and their values do not always linearly scale with factors
affecting fire behavior, such as fuel moisture. Providing estimates of fuel moisture
is thus of more value for operational, planning, and scientific purposes, as they
directly inform on a process that is relevant to fire behavior (but see Sect. 5.8) and
they can be directly used for modeling fire spread.
78 5 Environmental Plant Responses and Wildland Fire Danger

5.4 Drivers of Dead Fuel Moisture

Dead fuel is often divided into different categories depending upon the time lag to
reach equilibrium. For instance, 1 h fuel is that which takes less than 60 min to reach
63% (1−1/e) of the final value after exposure to new conditions. Using this approach,
dead fuel has been classified as 1 h, 10 h, 100 h, and 1000 h. In particles that are
suspended and exposed to free air, the time to reach equilibrium is dependent on the
diameter of the fuel particle. Consequently, 1 h, 10 h, 100 h, and 1000 h fuels are
often used to indicate particle thickness of < 6 mm, 6–25 mm, 25–75 mm, and
> 75 mm, respectively.
Not all dead fuels exert the same impact over fire behavior. Flame spread will be
mostly affected by the moisture status of small fuel particles (dead fine fuel).
Combustion of large fuel particles occurs after the flame has passed, and it will only
seldom contribute to fire behavior. Consequently, understanding the drivers of fine
fuel particles will be of utmost importance. Unfortunately, we lack the methods to
automatically measure 1 h fuel moisture. The sensors currently available to auto-
matically measure dead fuel use a pine dowel with a 10 h time lag. Available data on
1 h fuel thus depends upon manual measurements and are thus scarce. This method-
ological limitation has led to many studies focusing on understanding the drivers of
10 h fuel moisture.
Dead fuel moisture depends on three different processes: vapor exchange, latent
heat, and precipitation (Viney 1991). Vapor exchange leads to drying through the
evaporation from the fuel surface and through the desorption of water vapor. Vapor
exchange may also lead to wetting through adsorption. Wetting is additionally
affected by latent heat processes (such as condensation) and precipitation. Modeling
dead fuel moisture has encountered “an endless series of beautiful theories demol-
ished by ugly facts” (Chandler et al. 1983). Most of the research conducted to date
has focused on establishing dead fuel moisture models with high temporal resolu-
tion because models of fire behavior often require dead fuel moisture values at time
scales of minutes. The reader interested on this scale is referred to the different
reviews on this topic (Viney 1991; Matthews 2013).
Here we will focus on the landscape-scale drivers of dead fuel moisture.
Landscape predictions of dead fuel moisture are required for a large set of opera-
tions, such as developing maps of fire danger, determining whether environmental
conditions are suitable for a prescribed burn and modeling of fire activity. At this
scale, we can assume that evaporation will be the major driver of the drying rate of
fuel moisture. Moreover, for the purposes previously outlined, estimating fuel mois-
ture at the time of maximum dryness, that is, once per day, will generally suffice.
Considering all this, Resco de Dios et al. (2015) proposed that dead fine fuel
moisture depended largely on vapor pressure deficit. Following Fick’s law of diffu-
sion, evaporation will depend on the difference in vapor pressure between the fuel
and the atmosphere. Furthermore, for the purposes of dead fine fuel moisture mod-
eling, they demonstrated that fuel moisture could be approximated by air vapor
pressure deficit using a simple exponential decline function. Despite the large
5.5 Drivers of Live Fuel Moisture 79

i­ nfluence of dead fuel over fire activity, this is the only method that, to date, has been
tested with ground data to provide spatially explicit predictions of dead fuel
moisture.
This model may be expanded to provide regional estimates of fuel moisture
based on meteorological observations of temperature and relative humidity, if the
data are available (Nolan et al. 2016b). When such data are lacking, remotely sensed
land surface temperature may be used (Nolan et al. 2016b; Nieto et al. 2010). It
appears that dead fine fuel moisture estimated with this method is correlated with
cumulative burned area in contrasting parts of the world, such as the temperate for-
ests from SE Australia or from Portugal (Nolan et al. 2016a; Boer et al. 2017).
Interestingly, a dead fuel moisture content of 14% appears to be the seemingly
universal threshold of fuel dryness that, when crossed, may lead to catastrophic
wildfires in temperate forests (Boer et al. 2017). One of the advantages of predicting
dead fuel moisture with the approach presented in the paragraph above is that the
model does not require site-specific calibrations. The original model was calibrated
with data from a Cumberland Plain’s woodland near Sydney, and validations across
forests and shrublands in California and other parts of SE Australia indicated the
universality of the model (Nolan et al. 2016a).
Dead fuel moisture is just one of the components associated with fire danger.
However, under some circumstances, it may provide more realistic information on
fire danger than the indices currently in use. The Pedrógão Grande fire occurred in
Portugal in June 2017, burned ~45,000 ha, and led to 66 fatalities. That fire was
slightly off-season (wildland fires often occur after July in the area), and it followed
from an intense heat wave. Such heat wave led to a rapid decrease in dead fuel mois-
ture, unprecedented in the meteorological record. Indices used to predict fire danger
in the area (FWI) generally failed to recognize the extreme fire danger occurring
during those days as values were below critical thresholds. However, dead fuel
moisture was at 12%, well below the value for catastrophic fire (Boer et al. 2017).
This example highlights the need to move from fire indices to actual values of the
parameters that affect fire behavior and spread.
There are different types of wildland fires, which may be driven by topography,
wind, or fire plume, as we will discuss in more detail in Chap. 9. It is to be expected
that the relationship between dead fuel moisture content and area burned might be
different for each fire type as they respond differently to environmental variation.
However, we are still lacking more detailed assessments on whether different fuel
moisture thresholds act over different fire types.

5.5 Drivers of Live Fuel Moisture

The degree of variation in live fuel moisture content (LFMC) within a fire season
varies markedly across life-forms in Mediterranean environments (Fig. 5.1). LFMC
remains nearly constant in evergreen trees, but it can vary broadly in shrubs, particu-
larly in those lacking the capacity to resprout, and in herbs (Martin-StPaul et al.
80 5 Environmental Plant Responses and Wildland Fire Danger

(a) (b)
100

100
80

80
LFMC (%)

LFMC (%)
60

60
40

40
20

20
Non resprouter Resprouter Shrub Tree

(c)
140
120
LFMC (%)
100
80
60

160 180 200 220 240 260


day of year

Fig. 5.1 Seasonal variation in live fuel moisture content (ΔLFMC) across different species from
Southern France as function of post-fire regeneration strategy (a) and life-form (b). An example of
the seasonal pattern for a tree (Quercus ilex, continuous line) and a shrub (Genista scorpius, dashed
line) shown in (c). (Data from Martin-StPaul et al. (2018))

2018; Pellizzaro et al. 2007; Soler Martin et al. 2017; Viegas et al. 2001; Nolan et al.
2018; Yebra et al. 2019) (Fig. 5.1).
LFMC is defined as:

 F − Dw 
LFMC =  w ·100
 Dw  (5.1)
where Fw is the fresh weight of fuel and Dw is the dry weight of fuel. LFMC dynam-
ics are thus a function of variation in water mass and also in dry mass (Jolly and
Johnson 2018).
As explained by Nolan et al. (2018), there are three major physiological pro-
cesses affecting water mass. The first is access to water resources which, in turn,
depends on the interactions between soil water availability (or soil water potential)
5.5 Drivers of Live Fuel Moisture 81

and the rooting patterns. Rooting patterns appear to be related to the post-fire regen-
eration strategy. For example, species that regenerate from seed usually show shal-
lower root systems than those which can resprout (Bell and Pate 1996; Verdú 2000).
Consequently, the deeper rooting system observed in resprouting species may be
one of the reasons explaining why seasonal variation in LFMC is lower in resprout-
ers (Fig. 5.1). Water uptake is often a passive process, but it can be enhanced by the
symbiotic relationships with mycorrhizal fungi. That is, mycorrhizae might act as
an extended rooting system that allows for a larger exploration of the soil volume in
the search for water (Parke et al. 1983; Gehring et al. 2017). However, its impor-
tance in the field is not yet well established due to a paucity of data.
Once the water has been taken up by the rooting system and transported to the
twigs and leaves, there are two additional adjustments that may further affect
LFMC. One is the series of osmotic and elastic adjustments that take place in cells,
which lead to differences in turgor loss point and water storage capacities. That is,
as drought stress increases, the cells might (i) increase the elasticity of the cell walls
to avoid collapse under turgor loss (decrease of ε, the elasticity modulus); (ii)
increase the concentration of solutes in the vacuole to increase π, the osmotic poten-
tial at full turgor; or (iii) redistribute the symplastic water outside of the cell walls
towards apoplastic water (Bartlett et al. 2012). ε and π are both related to changes in
LFMC although the latter is considered to be more related to drought tolerance
(Bartlett et al. 2012).
The third physiological mechanism potentially explaining interspecific differ-
ences in leaf moisture is stomatal regulation and its effect over leaf water potential
(Ψleaf). Ψleaf is largely controlled by stomatal sensitivity to water stress. Traditionally,
plants exhibiting little seasonal variation in midday water potential have been clas-
sified as isohydric, while those with large fluctuations in water potential have been
considered anisohydric (Klein 2014). Differences in Ψleaf regulation strategies are
generally attributed to differing degrees of stomatal regulation (Martínez-Vilalta
and Garcia-Forner 2017; Martin-StPaul et al. 2017). Following stomatal closure,
Ψleaf can continue to decline due to stomatal leakiness and cuticular conductance
(Blackman et al. 2016). The relationship between Ψleaf and relative water content is
curvilinear (Scholander et al. 1965), and relative water content, in turn, shows a
high relationship with LFMC. The relationship between Ψleaf and relative water con-
tent (and thus LFMC) can further change through time due to osmotic or elastic
adjustments.
Apart from the physiological mechanisms leading to variation in water mass,
LFMC may also be affected by changes in dry mass (Jolly and Johnson 2018; Jolly
et al. 2014). Dry mass varies as a result of changes in nonstructural compounds
(e.g., nonstructural carbohydrates, which might vary within a diurnal cycle) and in
leaf mass area (LMA). LMA represents the investment in dry matter per area of
light-intercepting leaf tissue. A larger mass for a given area may require a larger
amount of water to maintain a given LFMC, and differences in LMA will also affect
the saturated water content. LMA is affected by leaf aging. Fresh tissues often show
low LMA, and in those species producing new tissue at the middle or end of the
summer, LFMC might increase as a result of new tissue production. LMA increases
82 5 Environmental Plant Responses and Wildland Fire Danger

a b

Fig. 5.2 Changes in leaf fuel moisture content in P. halepensis with age (a) and annual rainfall (b)

slowly over a season with leaf season, which has been linked with seasonal declines
in LFMC in conifers (Jolly et al. 2016). Furthermore, older evergreen leaves are
often drier than current-year leaves (Fig. 5.2), which is also a result of this process.
Usually plants show a coordinated response to water scarcity such that rooting
patterns, stomatal behavior, and osmotic and elastic adjustments all covary as envi-
ronmental stress increases (Nolan et al. 2018). From the point of view of fire danger
and spread, maybe we do not require a full understanding of the processes driving
the variation in LFMC. Instead, we can focus on selecting key traits that will allow
for mechanistic, simple, and accurate modeling of LFMC. A recent study indicated
that predawn water potential was correlated with LFMC. Predawn water potential is
an indicator of water availability in the rhizosphere, and it explains 63% of the
variation in LFMC across species from different life-forms (trees and shrubs) and
contrasting post-fire regeneration strategies (resprouters and seeders) in a
Mediterranean shrubland (Nolan et al. 2018). Measuring predawn water potential is
advantageous over measurements of LFMC because it is less time-consuming and
values may be obtained within minutes (unlike LFMC, where one needs to wait for
a few days to obtain dry weight). From the perspective of modeling, predawn water
potential is also advantageous as it is one of the simplest physiological parameters
to model. At the time of writing, there is no model of predawn water potential that
may be used for operational purposes at landscape scale, so this should be at the
forefront of our research efforts.
LFMC is currently being estimated through different remote sensing approaches
(Yebra et al. 2013, 2018). This is because the reflectance of plant tissues varies with
water content. These approaches allow for monitoring of past or current conditions
but do not allow forecasting. Forecasting may be achieved by inferring LFMC from
meteorological drought indices, but the success of this practice seems to be limited
(Ruffault et al. 2018; Soler Martin et al. 2017). Coupling physiological models of
LFMC with remotely sensed estimates of LFMC thus shows a large potential to
overcome current limitations to develop spatially explicit predictions of LFMC, but
the challenge lies in developing robust physiological models that operate at regional
5.6 Problems in Measuring and Estimating Fuel Moisture in Plant Canopies 83

scales. Furthermore, spatially explicit LFMC forecasts could be better constrained


with remote sensing via data assimilation.

5.6  roblems in Measuring and Estimating Fuel Moisture


P
in Plant Canopies

Measurements of canopy moisture are often performed manually by weighing fresh


and oven-dried fine twigs and leaves. These measurements are performed for scien-
tific purposes, and some fire agencies also routinely estimate canopy moisture. An
incomplete understanding of the physiology of fuel moisture may lead to biased
estimates. Within-canopy moisture is spatially heterogeneous, but current sampling
efforts provide a single value of fuel moisture per species or per fuel strata (depend-
ing on the sampling design). There are different processes leading to spatial hetero-
geneity in canopy moisture, as described in the previous section. Differences across
species are often accounted for by sampling individually the different species or by
taking a composite sample considering the relative proportion of the different co-­
occurring species.
Differences in canopy moisture within a species, and even within a canopy, are
more difficult to correct for because leaf and shoot age create further heterogeneity.
For instance, if only current-year shoots are sampled, fuel moisture will be overes-
timated (Chrosciewicz 1986). Any appropriate sampling design needs to take into
account leaf lifespan and sample across all leaf ages (Fig. 5.2).
The major problem in the existence of spatial variation in canopy moisture how-
ever comes from the existence of dead fuels in plant canopies (Fig. 5.3). Indeed,
plant canopies are composed of live and of senesced tissue, and the proportion var-
ies within a fire season (see Chap. 6). Manual measurements of fuel moisture should
thus incorporate this process.
In addition to manual measurements, remote sensing is often used to estimate
live fuel moisture content at a regional scale. Unfortunately, satellites often lack the
appropriate spatial resolution to account, in particular, for the proportion of dead
biomass and its moisture value. That is, satellite-derived estimates of LFMC pro-
vide an averaged value of moisture in the canopy that integrates live and dead mat-
ter. However, we will not be able to distinguish the % of the signal originating from
dead or from live biomass. Estimates of % dead biomass, and their effect over can-
opy foliage, may be obtained through different sensors mounted at either fixed plat-
forms or aerial platforms such as UAVs (Santini et al. 2019). Phenocams provide the
most inexpensive solution. Phenocams allow to quantify the proportion of dead
biomass in grass fuels by, for example, examining the green chromatic coordinate
index (Collins et al. 2018; Sonnentag et al. 2012). GCC is calculated from the pro-
portion of green within a red-green-blue (RGB) image. Within forest canopies,
GCC has been used to estimate phenological events, and estimates of the percent
84 5 Environmental Plant Responses and Wildland Fire Danger

Fig. 5.3 Example of pre-programmed needle senescence in a Mediterranean conifer (a) and of
drought-induced senescence (b) in a temperate eucalypt woodland. Both lead to a substantial
increase in canopy dead matter, thus lowering canopy fuel moisture and enhancing crown fire
likelihood. However, while only ~ 1/3 of the canopy is dry in the conifer forest (needle lifespan in
the P. halepensis in the picture is 3 years, and only older leaves are shed by pre-programmed cell
death), the entire canopy is dry in the eucalypt forest. The danger of crown fire is thus higher in (b)
than in (a). This increase in fire danger is only transient as canopy fuels are deposited on the
ground after a few weeks, lowering the danger of crown fire. Pre-programmed needle senescence
refers to a developmental ontogenetic process, whereas drought-induced senescence may be driven
by hydraulic failure or other processes, as we discuss more fully in Chap. 6. (Photo credits: Carles
Arteaga (a) and Rachael H Nolan (b))

dead foliage will be more complicated than in grasslands because a large portion of
the image will be wood.
There are indeed a series of methodological issues that need to be taken into
account because GCC is affected by illumination and shadow (Sonnentag et al.
2012). For instance, it is recommended to take photographs under anisotropic light
conditions, such as those occurring in the early morning. Full details on method-
ological considerations are available elsewhere (Sonnentag et al. 2012; Toomey
et al. 2015; Migliavacca et al. 2011; Collins et al. 2018).
After quantification of the proportion of dead canopy, we still need to understand
how this value will affect the moisture content of the canopy and whether it will
affect fire behavior. As mentioned in the previous chapter, Pinus halepensis leaves
live for 3 years, and leaf shedding of the older leaves occurs either at the end of June
or at the start of July, depending on the year, which coincides with the onset of the
fire season in the Western Mediterranean Basin. Anecdotal evidence indicates that
fires burning pine forests at the time of leaf senescence exhibit higher intensities
(M. Castellnou, pers. comm.), but proper quantification of this issue, and its poten-
tial incorporation in fire behavior models, is still awaiting.
5.7 Fuel Moisture and Fire Occurrence 85

5.7 Fuel Moisture and Fire Occurrence

The role of dead fuel moisture in affecting fire spread is well known. There is how-
ever an ongoing debate as to whether live fuel moisture also affects fire spread.
LFMC should affect fire spread because high moisture values should increase the
energy required for the fire to propagate. That is, the energy incoming to the fuel
should first be transformed into latent heat (evaporation), and ignition would occur
only once the fuel is dry (hot) enough to burn. There is some correlative evidence
indicating a link between LFMC and fire activity. Studies conducted in SE Australia,
Spain, California, and China all coincide that fires only occur after a threshold value
of fuel dryness has been crossed (Jurdao and Chuvieco 2012; Dennison et al. 2008;
Nolan et al. 2016a; Luo et al. 2019). This threshold value appears to vary with eco-
systems and the tree-to-shrub ratio. In the temperate or subtropical forest ecosys-
tems from SE Australia, fire activity increases markedly once LFMC drops below
ca. 100% (Fig. 5.4b). However in the Mediterranean ecosystems from California,
where a higher proportion of shrubs occur, this threshold drops to ca. 70% (Fig. 5.4a).
Further evidence on the relevance of LFMC for fire spread comes from experi-
mental studies conducted in the laboratory, where relationships between LFMC and
rate of spread are also found (Alexander and Cruz 2013). However, and in contrast
with laboratory-based studies, a review of the literature on individual fires indicated
that no field study has so far observed a significant relationship between the rate of
fire spread and LFMC (Alexander and Cruz 2013). This literature review supported
the notion that wildland fires occur once a threshold value of moisture is crossed but
concluded that further decreases in LFMC do not affect the rate of spread. They
hypothesized that the discrepancy between laboratory and field results lies in the

a b
Cumulative burned area (km2)
Cumulative burned area (ha)

30000
30000

10000
10000
0

60 80 100 120 140 60 80 100 120 140 160


LFMC (%) LFMC (%)

Fig. 5.4 Sharp increases in burned area are observed below certain thresholds of live fuel moisture
area in California (a) and Australia (b). The lines indicate the results of segmented regression and
were calculated after digitizing data from Dennison et al. (2008) and Nolan et al. (2016a)
86 5 Environmental Plant Responses and Wildland Fire Danger

different amounts of energy released in each fire type. Fires in the laboratory show
a generally small radiative power, compared with field fires, and thus the effect of
LFMC may be exaggerated. The large intensity released by fires in the field would
provide enough heat to ignite the fuel before all the water is vaporized.
Some authors have tried to explain this discrepancy between laboratory and field
studies based upon the differential temporal patterns of moisture content across life-­
forms and fuel types (Rossa and Fernandes 2018a, b). More specifically, it has been
proposed that relationships between LFMC and fire activity are confounded because
of the different patterns of LFMC across trees and shrubs. Whereas LFMC in trees
is nearly constant, significant declines are often observed in shrubs during the sum-
mer (Figs. 5.1 and 5.5) as previously described. Moreover, these authors state that
the seasonal pattern of moisture in shrubs is correlated with the seasonal pattern of
dead fuel moisture (Rossa and Fernandes 2018a). The differential moisture pattern
across life-forms, and a potential correlation between shrub and dead fuel moisture,
would thus mask any potential effect of LFMC over fire spread.
However, this hypothesis is not exempt of criticism. Based upon the information
we have provided in Sects. 5.4 and 5.5, we would expect the temporal pattern
in shrub LFMC to be decoupled from that in dead fuel moisture content (DFMC).
Shrub LFMC should exhibit a monotonic decline as the drought intensity increases
in those areas that experience long and dry summers, such as the Mediterranean
Basin. In contrast, DFMC should remain stable and low through the fire season. We
have previously indicated that critical thresholds for fire activity occur when DFMC
drops below 14%, and, following the model from Resco de Dios et al. (2015), this
happens under a vapor pressure deficit of 2.7 kPa. In the long and dry Mediterranean

a b

Fig. 5.5 The temporal patterns of long-term (20 years) variation in live (LFMC) and dead (DMFC)
fuel moisture content (a) are not significantly correlated with each other (b). Data originates from
Rossa and Fernandes (2018b), who provided data for a shrubland with Calluna vulgaris and
Pterospartum tridentatum. Neither correlations between LFMC and DFMC were statistically sig-
nificant in (b) when comparing the data for the entire year (P = 0.54) or when considering only data
within the fire season (P = 0.29)
5.8 Relationships Between Drought and Fire Occurrence 87

summers, this implies that DFMC will be low and below the critical flammability
threshold for much of the fire season. There is thus little reason to expect a correla-
tion between shrub LFMC, which should monotonically decline over the summer,
and DFMC, which should remain reasonably stable and at low values during
that time.

5.8 Relationships Between Drought and Fire Occurrence

A consequence of the potential relationship between LFMC and the rate of spread
is that increases in the duration of water scarcity (drought) should be accompanied
by increases in burned area. Increasing drought leads to declines in LFMC, and, if
LFMC affects fire spread, we should observe a higher area burned as drought
increases. We can thus examine the temporal pattern of fire and drought occurrences
to obtain further insights on the relationship between LFMC and fire activity. Fire
activity in the Mediterranean part of Spain peaks in the first 2 weeks of July
(Fig. 5.6). Based upon the dynamics of shrub LFMC, which mirror drought dura-
tion, we should expect fire activity to peak at the driest part of the summer, which is
towards the end of the summer in late August. However, burned area declines as the
summer advances (Fig. 5.6), while LFMC in shrubs also declines (Figs. 5.1 and
300000
Burned area (ha)
200000
100000
0

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Fig. 5.6 Temporal pattern of long-term average (1968–2015) burned area in the Mediterranean
part of Spain. (Data from the Estadística General de Incendios Forestal provided by the Ministry
of Agriculture, Fishing and Food. Each value indicates fortnightly averages)
88 5 Environmental Plant Responses and Wildland Fire Danger

5.5). This simple analysis does not unequivocally preclude a LFMC effect on fire
intensity. Furthermore, burned area peaks at different times in different parts of the
globe. However, this simple analysis indicates that, within Mediterranean ecosys-
tems, additional factors apart from LFMC also need to be taken into account to
explain the temporal pattern of burned area.
To conclude, it is clear that there is enough evidence to indicate that moisture
levels in the fuel complex need to be beyond critical flammability thresholds: 14%
for dead and 70–100% for live fuels. This means that fire-prone ecosystems remain
in a critical state for much of the year. However, once the ecosystem has crossed this
critical flammability threshold, that is, once it is dry enough to burn, it is not clear
that further increases in drought, which mainly impact LFMC, will exert any further
influence on fire activity. The commonly assumed hypothesis that more drought
leads to more fire thus needs to be examined in further detail. Understanding the
mechanisms affecting fuel moisture content and its relationship with fire activity,
that is, the pyrophysiology of environmental fire danger, should be at the forefront
of our research efforts. In particular, interactions between fuel moisture and changes
in synoptic conditions that occur during the fire season (Duane and Brotons 2018;
Potter 2012; Peterson et al. 2017; Lewis et al. 2019) might help to better explain
seasonal changes in fire danger and burned area.

References

Abatzoglou JT, Kolden CA (2013) Relationships between climate and macroscale area burned
in the western United States. Int J Wildland Fire 22(7):1003–1020. https://doi.org/10.1071/
WF13019
Aguado I, Chuvieco E, Borén R, Nieto H (2007) Estimation of dead fuel moisture content from
meteorological data in Mediterranean areas. Applications in fire danger assessment. Int J
Wildland Fire 16(4):390–397. https://doi.org/10.1071/WF06136
Alexander ME, Cruz MG (2013) Assessing the effect of foliar moisture on the spread rate of crown
fires. Int J Wildland Fire 22:415–427. https://doi.org/10.1071/wf12008_co
Bartlett MK, Scoffoni C, Sack L (2012) The determinants of leaf turgor loss point and prediction
of drought tolerance of species and biomes: a global meta-analysis. Ecol Lett 15(5):393–405.
https://doi.org/10.1111/j.1461-0248.2012.01751.x
Bedia J, Herrera S, Gutiérrez JM, Benali A, Brands S, Mota B, Moreno JM (2015) Global patterns
in the sensitivity of burned area to fire-weather: implications for climate change. Agric For
Meteorol 214-215:369–379. https://doi.org/10.1016/j.agrformet.2015.09.002
Bell TL, Pate JS (1996) Growth and fire response of selected Epacridaceae of south- western
Australia. Aust J Bot 44:509–526
Berry SL, Roderick ML (2005) Plant-water relations and the fibre saturation point. New Phytol
168(1):25–37. https://doi.org/10.1111/j.1469-8137.2005.01528.x
Blackman CJ, Pfautsch S, Choat B, Delzon S, Gleason SM, Duursma RA (2016) Toward an index
of desiccation time to tree mortality under drought. Plant Cell Environ 39:2342–2345. https://
doi.org/10.1111/pce.12758
Boer MM, Nolan RH, Resco De Dios V, Clarke H, Price OF, Bradstock RA (2017) Changing
weather extremes call for early warning of potential for catastrophic fire. Earth’s Future
5:1196–1202. https://doi.org/10.1002/2017EF000657
References 89

Boer MM, Resco de Dios V, Stefaniak E, Bradstock RA (2019) A hydroclimatic model for the
distribution of fire on Earth. Biogeosci Discuss
Bradstock RA (2010) A biogeographic model of fire regimes in Australia: current and future impli-
cations. Glob Ecol Biogeogr 19:145–158. https://doi.org/10.1111/j.1466-8238.2009.00512.x
Chandler C, Cheney P, Thomas P, Trabaud L, Williams DG (1983) Fire in forestry, vol 1. Forest
fire behavior and effects. Wiley, New York
Chrosciewicz Z (1986) Foliar moisture content variations in four coniferous tree species of central
Alberta. Can J For Res 16(1):157–162. https://doi.org/10.1139/x86-029
Collins L, Bradstock RA, Resco de Dios V, Duursma RA, Velasco S, Boer MM (2018) Understorey
productivity in temperate grassy woodland responds to soil water availability but not to ele-
vated [CO2]. Glob Chang Biol 24(6):2366–2376. https://doi.org/10.1111/gcb.14038
Dennison PE, Moritz MA, Taylor RS (2008) Evaluating predictive models of critical live fuel
moisture in the Santa Monica Mountains, California. Int J Wildland Fire 17:18–27
Duane A, Brotons L (2018) Synoptic weather conditions and changing fire regimes in a
Mediterranean environment. Agric For Meteorol 253-254:190–202. https://doi.org/10.1016/j.
agrformet.2018.02.014
Fernandes PM, Barros AMG, Pinto A, Santos JA (2016) Characteristics and controls of extremely
large wildfires in the western Mediterranean Basin. J Geophys Res Biogeo 121(8):2141–2157.
https://doi.org/10.1002/2016JG003389
Flannigan MD, Wotton BM, Marshall GA, de Groot WJ, Johnston J, Jurko N, Cantin AS (2016)
Fuel moisture sensitivity to temperature and precipitation: climate change implications. Clim
Chang 134(1):59–71. https://doi.org/10.1007/s10584-015-1521-0
Gehring CA, Sthultz CM, Flores-Rentería L, Whipple AV, Whitham TG (2017) Tree genetics
defines fungal partner communities that may confer drought tolerance. Proc Natl Acad Sci
114(42):11169–11174. https://doi.org/10.1073/pnas.1704022114
Jolly W, Johnson D (2018) Pyro-ecophysiology: shifting the paradigm of live wildland fuel
research. Fire 1(1):8. https://doi.org/10.3390/fire1010008
Jolly WM, Hadlow AM, Huguet K (2014) De-coupling seasonal changes in water content and dry
matter to predict live conifer foliar moisture content. Int J Wildland Fire 23(4):480–489. https://
doi.org/10.1071/wf13127
Jolly WM, Hintz J, Linn RL, Kropp RC, Conrad ET, Parsons RA, Winterkamp J (2016) Seasonal
variations in red pine (Pinus resinosa) and jack pine (Pinus banksiana) foliar physio-chemistry
and their potential influence on stand-scale wildland fire behavior. For Ecol Manag 373:167–
178. https://doi.org/10.1016/j.foreco.2016.04.005
Jurdao S, Chuvieco E (2012) Modelling fire ignition probability from satellite estimates of live fuel
moisture content. Fire Ecol 7:77–97. https://doi.org/10.4996/fireecology.0801077
Karavani A, Boer MM, Baudena M, Colinas C, Díaz-Sierra R, Pemán J, de Luís M, Enríquez-­
de-­Salamanca Á, Resco de Dios V (2018) Fire-induced deforestation in drought-prone
Mediterranean forests: drivers and unknowns from leaves to communities. Ecol Monogr
88:141–169
Keetch JJ, Byram G (1968) A drought index for forest fire control. Research Paper SE-38.
U.S. Department of Agriculture, Forest Service, Southeastern Forest Experiment Station,
Asheville
Klein T (2014) The variability of stomatal sensitivity to leaf water potential across tree species
indicates a continuum between isohydric and anisohydric behaviours. Funct Ecol 28:1313–
1320. https://doi.org/10.1111/1365-2435.12289
Lewis SC, Blake SAP, Trewin B, Black MT, Dowdy AJ, Perkins-Kirkpatrick SE, King AD,
Sharples JJ (2019) Deconstructing factors contributing to the 2018 fire weather in Queensland,
Australia. B Am Meteorol Soc. https://doi.org/10.1175/BAMS-D-19-0144.1
Luo K, Quan X, He B, Yebra M (2019) Effects of live fuel moisture content on wildfire occur-
rence in fire-prone regions over Southwest China. Forests 10(10):887. https://doi.org/10.3390/
f10100887
90 5 Environmental Plant Responses and Wildland Fire Danger

Macias Fauria M, Michaletz ST, Johnson EA (2011) Predicting climate change effects on wild-
fires requires linking processes across scales. Wiley Interdiscip Rev Clim Chang 2(1):99–112.
https://doi.org/10.1002/wcc.92
Martínez-Vilalta J, Garcia-Forner N (2017) Water potential regulation, stomatal behaviour and
hydraulic transport under drought: deconstructing the iso/anisohydric concept. Plant Cell
Environ 40(6):962–976. https://doi.org/10.1111/pce.12846
Martin-StPaul N, Delzon S, Cochard H (2017) Plant resistance to drought depends on timely sto-
matal closure. Ecol Lett 20(11):1437–1447. https://doi.org/10.1111/ele.12851
Martin-StPaul N, Pimont F, Dupuy JL, Rigolot E, Ruffault J, Fargeon H, Cabane E, Duché Y,
Savazzi R, Toutchkov M (2018) Live fuel moisture content (LFMC) time series for multiple
sites and species in the French Mediterranean area since 1996. Ann For Sci 75(2):57. https://
doi.org/10.1007/s13595-018-0729-3
Matthews S (2013) Dead fuel moisture research: 1991–2012. Int J Wildland Fire 23:78–92
McArthur AG (1967) Fire behaviour in eucalypt forest. Comm Aust For Timb Bur Leaflet 107:25pp
Migliavacca M, Galvagno M, Cremonese E, Rossini M, Meroni M, Sonnentag O, Cogliati S, Manca
G, Diotri F, Busetto L, Cescatti A, Colombo R, Fava F, Morra di Cella U, Pari E, Siniscalco C,
Richardson AD (2011) Using digital repeat photography and eddy covariance data to model
grassland phenology and photosynthetic CO2 uptake. Agric For Meteorol 151(10):1325–1337.
https://doi.org/10.1016/j.agrformet.2011.05.012
Nelson RF (2001) Water relations of forest fuels. In: Johnson EA, Miyanishi K (eds) Forest fires:
behavior and ecological effects. Academic Press, New York, pp 79–149
Nieto H, Aguado I, Chuvieco E, Sandholt I (2010) Dead fuel moisture estimation with MSG–SEVIRI
data. Retrieval of meteorological data for the calculation of the equilibrium moisture content.
Agric For Meteorol 150(7–8):861–870. https://doi.org/10.1016/j.agrformet.2010.02.007
Nogueira J, Rambal S, Barbosa J, Mouillot F (2017) Spatial pattern of the seasonal drought/burned
area relationship across Brazilian biomes: sensitivity to drought metrics and global remote-­
sensing fire products. Climate 5(2):42
Nolan RH, Boer MM, Resco de Dios V, Caccamo G, Bradstock RA (2016a) Large scale, dynamic
transformations in fuel moisture drive wildfire activity across south-eastern Australia. Geophys
Res Lett 43:4229–4238
Nolan RH, Resco de Dios V, Boer MM, Caccamo G, Goulden ML, Bradstock RA (2016b)
Predicting dead fine fuel moisture at regional scales using vapour pressure deficit from MODIS
and gridded weather data. Remote Sens Environ 174:100–108. https://doi.org/10.1016/j.
rse.2015.12.010
Nolan RH, Hedo J, Arteaga C, Sugai T, Resco de Dios V (2018) Physiological drought responses
improve predictions of live fuel moisture dynamics in a Mediterranean forest. Agric For
Meteorol 263:417–427. https://doi.org/10.1016/j.agrformet.2018.09.011
Parke JL, Linderman RG, Black CH (1983) The role of ectomycorrhizas in drought tolerance of
douglas-fir seedlings. New Phytol 95:83–95
Pellizzaro G, Duce P, Ventura A, Zara P (2007) Seasonal variations of live moisture content and
ignitability in shrubs of the Mediterranean Basin. Int J Wildland Fire 16:633–641
Peterson DA, Hyer EJ, Campbell JR, Solbrig JE, Fromm MD (2017) A conceptual model for
development of intense pyrocumulonimbus in western north america. Mon Weather Rev
145(6):2235–2255. https://doi.org/10.1175/mwr-d-16-0232.1
Potter BE (2012) Atmospheric interactions with wildland fire behaviour – I. Basic surface inter-
actions, vertical profiles and synoptic structures. Int J Wildland Fire 21(7):779. https://doi.
org/10.1071/wf11128
Resco de Dios V, Fellows AW, Nolan RH, Boer MM, Bradstock RA, Domingo F, Goulden ML
(2015) A semi-mechanistic model for predicting the moisture content of fine litter. Agric For
Meteorol 203:64–73
Rossa C, Fernandes P (2018a) Live fuel moisture content: the ‘pea under the mattress’ of fire
spread rate modeling? Fire 1(3):43
References 91

Rossa CG, Fernandes PM (2018b) Short communication: on the effect of live fuel moisture content
on fire-spread rate. For Syst 26(3):eSC08. https://doi.org/10.5424/fs/2017263-12019
Ruffault J, Martin-StPaul N, Pimont F, Dupuy J-L (2018) How well do meteorological drought
indices predict live fuel moisture content (LFMC)? An assessment for wildfire research
and operations in Mediterranean ecosystems. Agric For Meteorol 262:391–401. https://doi.
org/10.1016/j.agrformet.2018.07.031
Santini F, Kefauver SC, Dios VRd, Araus JL, Voltas J (2019) Imaging spectroscopy for high-­
throughput phenotyping of forest genetic trials: a case study in Pinus halepensis. Ann Apl Bot
(in press)
Scholander PF, Hammel HT, Bradstreet ED, Hemmingsen EA (1965) Sap pressure in vascular
plants. Science 148(3668):339–346
Schunk C, Wastl C, Leuchner M, Menzel A (2017) Fine fuel moisture for site- and species-specific
fire danger assessment in comparison to fire danger indices. Agric For Meteorol 234-235:31–
47. https://doi.org/10.1016/j.agrformet.2016.12.007
Sharples JJ, McRae RHD, Weber RO, Gill AM (2009) A simple index for assessing fire danger rat-
ing. Environ Model Softw 24(6):764–774. https://doi.org/10.1016/j.envsoft.2008.11.004
Soler Martin M, Bonet JA, Martínez De Aragón J, Voltas J, Coll L, Resco De Dios V (2017) Crown
bulk density and fuel moisture dynamics in Pinus pinaster stands are neither modified by thin-
ning nor captured by the Forest Fire Weather Index. Ann For Sci 74:51. https://doi.org/10.1007/
s13595-017-0650-1
Sonnentag O, Hufkens K, Teshera-Sterne C, Young AM, Friedl M, Braswell BH, Milliman T,
O’Keefe J, Richardson AD (2012) Digital repeat photography for phenological research
in forest ecosystems. Agric For Meteorol 152 (0):159–177. doi:https://doi.org/10.1016/j.
agrformet.2011.09.009
Sullivan AL (2009) Wildland surface fire spread modelling, 1990–2007. 2: empirical and quasi-­
empirical models. Int J Wildland Fire 18:369–386
Tian X, McRae DJ, Jin J, Shu L, Zhao F, Wang M (2011) Wildfires and the Canadian Forest Fire
Weather Index system for the Daxing’anling region of China. Int J Wildland Fire 20(8):963–
973. https://doi.org/10.1071/WF09120
Toomey M, Friedl MA, Frolking S, Hufkens K, Klosterman S, Sonnentag O, Baldocchi DD,
Bernacchi CJ, Biraud SC, Bohrer G, Brzostek E, Burns SP, Coursolle C, Hollinger DY,
Margolis HA, McCaughey H, Monson RK, Munger JW, Pallardy S, Phillips RP, Torn MS,
Wharton S, Zeri M, Richardson AD (2015) Greenness indices from digital cameras predict
the timing and seasonal dynamics of canopy-scale photosynthesis. Ecol Appl 25(1):99–115.
doi:doi:https://doi.org/10.1890/14-0005.1
Urbieta IR, Zavala G, Bedia J, Gutiérrez JM, Miguel-Ayanz JS, Camia A, Keeley JE, Moreno
JM (2015) Fire activity as a function of fire-weather seasonal severity and antecedent climate
across spatial scales in southern Europe and Pacific western USA. Env Res Lett 10:114013.
https://doi.org/10.1088/1748-9326/10/11/114013
van Wagner CE (1987) Development and structure of the Canadian Forest Fire Weather Index
system. Canadian Forestry Service, Ottawa, ON
Verdú M (2000) Ecological and evolutionary differences between Mediterranean seeders and
resprouters. J Veg Sci 11(2):265–268. https://doi.org/10.2307/3236806
Viegas DX, Piñol J, Viegas MT, Ogaya R (2001) Estimating live fine fuels moisture content using
meteorologically-based indices. Int J Wildland Fire 10:223–240
Viney N (1991) A review of fine fuel moisture modelling. Int J Wildland Fire 1:215–234. https://
doi.org/10.1071/WF9910215
Vinodkumar, Dharssi I, Bally J, Steinle P, McJannet D, Walker J (2017) Comparison of soil wet-
ness from multiple models over Australia with observations. Water Resour Res 53(1):633–646.
https://doi.org/10.1002/2015WR017738
Watson A, Lovelock JE, Margulis L (1978) Methanogenesis, fires and the regulation of atmo-
spheric oxygen. Biosystems 10(4):293–298. https://doi.org/10.1016/0303-2647(78)90012-6
92 5 Environmental Plant Responses and Wildland Fire Danger

Wotton BM, Beverly JL (2007) Stand-specific litter moisture content calibrations for the Canadian
Fine Fuel Moisture Code. Int J Wildland Fire 16(4):463–472. https://doi.org/10.1071/WF06087
Yebra M, Dennison PE, Chuvieco E, Riaño D, Zylstra P, Hunt ER Jr, Danson FM, Qi Y, Jurdao S
(2013) A global review of remote sensing of live fuel moisture content for fire danger assess-
ment: moving towards operational products. Remote Sens Environ 136:455–468. https://doi.
org/10.1016/j.rse.2013.05.029
Yebra M, Quan X, Riaño D, Rozas Larraondo P, van Dijk AIJM, Cary GJ (2018) A fuel moisture
content and flammability monitoring methodology for continental Australia based on optical
remote sensing. Remote Sens Environ 212:260–272. https://doi.org/10.1016/j.rse.2018.04.053
Yebra M, Scortechini G, Badi A, Beget ME, Boer MM, Bradstock R, Chuvieco E, Danson FM,
Dennison P, Resco de Dios V, Di Bella CM, Forsyth G, Frost P, Garcia M, Hamdi A, He B,
Jolly M, Kraaij T, Martín MP, Mouillot F, Newnham G, Nolan RH, Pellizzaro G, Qi Y, Quan
X, Riaño D, Roberts D, Sow M, Ustin S (2019) Globe-LFMC, a global plant water status data-
base for vegetation ecophysiology and wildfire applications. Scient Data 6:155. ­doi:https://doi.
org/10.1038/s41597-019-0164-9
Chapter 6
Plant Carbon Economies
and the Dynamics of Wildland Fuels

Abstract Fuel is anything that can burn. In a wildfire, fuel will be largely com-
posed of either plant tissues or dead plant matter. There are marked differences in
growth forms between plants and in the functional attributes that occur within and
across growth forms. Plant diversity leads to differences in fuel composition and
structure, and consequently, it may exert cascading effects on fire behavior. Fire
scientists and managers are well aware of the different structural attributes that may
affect behavior, but understanding why structural differences occur across plant spe-
cies or plant functional types may be less well understood. In this chapter, we will
describe plant growth and seasonal trait dynamics in order to understand the dynam-
ics of wildland fuel formation and its variation across vegetation types.

6.1 Wildland Fuels and Fire

Wildland fuels are plants, both dead and alive, and often vary from seasonal to
annual scales. Annual or decadal variation in wildland fuel loads occurs as the eco-
system is developing after disturbance (Fig. 6.1). In this chapter we will develop a
general understanding on how the structural and chemical properties that affect
wildfire behavior originate (Table 6.1).
We begin by explaining the development of live wildland fuels from the perspec-
tive of carbon and biomass accumulation patterns. We briefly explain the origin of
the different life-forms we encounter in fire-prone ecosystems, how wildland fuels
grow and develop, and how differences in crown architecture, essential for crown
fire behavior, arise. This discussion occurs within a framework of post-fire regen-
eration strategies, ultimately describing processes that lead to seasonal changes in
fuel flammability, including volatiles, inorganic chemistry, and leaf anatomy. An
obvious process that changes seasonally is fuel moisture. This chapter is focused on
biomass dynamics, and fuel moisture has already been discussed in Chap. 5. We
will then discuss the development of dead wildland fuels, including leaf shedding,
self-pruning and self-thinning, bark shedding, and plant death. Finally, we will dis-
cuss deposition and the decomposition of dead fuels (Fig. 6.1).

© Springer Nature Switzerland AG 2020 93


V. Resco de Dios, Plant-Fire Interactions, Managing Forest Ecosystems 36,
https://doi.org/10.1007/978-3-030-41192-3_6
94 6 Plant Carbon Economies and the Dynamics of Wildland Fuels

Live fuels
Ecosystem Development

Growth
Allocation Canopy fuels
Anatomy (structure, chemistry, anatomy)
VOC
Chemistry
Canopy Phenology Deposition
Firebrands

Seasonal Years
Surface fuels
Dead fuels
Ecosystem
Decomposition
Mortality
Self-pruning
Ground fuels
Bark shedding
Canopy Senescence

Seasonal Years

Fig. 6.1 Physiological development of live and dead fuels and how they affect crown, surface, and
ground fuel dynamics through deposition and decomposition. Senescence and deposition play a
key role in increasing dead matter in canopies, which then significantly affects the likelihood and
spread rate of a crown fire

6.2 Development of Live Wildland Fuels

6.2.1 Different Life-Forms

Wildland fuels are composed of (dead or alive) plants. The vast diversity of plant
species that exist on Earth, along with the dramatic differences in the functional
traits that they exhibit, result in substantial differences in the structure and organiza-
tion of wildland fuels. The major observable difference across live fuel types is the
different growth forms. Within an ecosystem we may observe trees, shrubs, and
herbs, to name a few. Plants grow, and consequently fuels develop, from apical and
lateral meristematic tissues, which have the capacity to divide. One of the possible
ways to divide different life-forms derives from the differences in the location of
buds, which is a particular type of meristematic tissue. Differences in bud location
are thus a major contributing factor to the occurrence of different strata of live fuel
within an ecosystem.
The initial classification of life-forms depending on the location of buds was
provided by Raunkiaer (1934). It has been expanded substantially in subsequent
work by different authors, but here we follow one of the simplest and most widely
6.2 Development of Live Wildland Fuels 95

Table 6.1 Traits affecting different aspects of the fire regime


Fuel trait Driver Major fire effects
Fire type
Life-form Bud location Fuel structure (grassy,
forest, etc.)
Fuel loading across Growth rates, phenology, deposition, Fire intensity
crown, surface, and decomposition
ground
Ladder fuels Life-form diversity, self-pruning, growth Probability of crown fire
rates
Live/dead Phenology, curing rates, self-thinning,
disturbance
Fire intensity and behavior
Surface fires
Biomass and depth Growth rates, deposition Fire intensity and rate of
spread
Heat content Chemical composition Fire intensity
Surface area to Leaf mass area, leaf shape and size Probability of ignition,
volume ratio and time-to-ignition
diameter
Mineral content Tissue type, physiology Reaction intensity
Fuel moisture Leaf anatomy and physiology, rooting Reaction intensity
depth structure
Packing ratio Leaf mass area, leaf shape, and size Rate of spread
Live/dead Phenology, curing rates, self-thinning, Rate of spread
disturbance
Crown fires
Canopy bulk density Canopy architecture, light interception, Rate of spread
Canopy depth optimization of carbon gain, canopy and
Canopy width aerodynamic conductance, growth rates
Canopy base height Self-pruning, growth rates Likelihood of crown fire
Foliage moisture Leaf anatomy and physiology, rooting Likelihood of crown fire
depth structure and spread
Live/dead ratio Phenology, curing rates, self-thinning, Rate of spread
disturbance
Foliage and branch Canopy architecture, light interception, Fire intensity
biomass optimization of carbon gain, canopy and
aerodynamic conductance, growth rates,
deposition
Self-thinning Biomass accumulation Likelihood of crown fire
Fire season
Leaf moisture below Leaf anatomy and physiology, rooting Fires only occur below
threshold values depth structure certain moisture thresholds
Fuel accumulation Phenology, curing rates, self-thinning, Fires only occur after
disturbance certain fuel loads
Fire return interval
(continued)
96 6 Plant Carbon Economies and the Dynamics of Wildland Fuels

Table 6.1 (continued)


Fuel trait Driver Major fire effects
Life-form Bud location Affect regrowth structure
Growth rates Efficiency in resources uptake and Determines the time
allocation necessary to build up
enough fuel to carry
another fire
Mosaic
Fuel spatial structure Ecological processes driving differences in Determine the degree of
land cover spatial connectivity
Firebrand availability Reproductive structure (e.g., cones vs Determines the potential
acorns), decorticating bark, leaf structure spotting distance
(needle vs broadleaf)
The table highlights the major fire effects of different fuel traits and its drivers
Note that additional fire effects not noted here are possible. The section on fire intensity is largely
based on a single model (Rothermel 1972; Albini 1976), and alternative views are also possible.
Regardless of the model used, these traits are usually considered as major drivers of fire intensity
and behavior

used schemes for forest plant species (note that we will not include non-forest
species).
1. Phanerophytes include trees and shrubs that grow taller than 50 cm and do not
experience periodic dieback to that height. Their buds may be found high above
the ground on aerial shoots.
2. Chamaephytes show a shoot or branch system that does not grow above 0.5 m,
or, if it does, the shoots die back periodically to 50 cm or below (e.g., dwarf
shrubs). Buds in chamaephytes are above the ground, but close to it, typically
below 25 cm in height.
3. Hemicryptophytes are perennial grasses or rosette forbs with buds at the ground
surface.
4. Geophytes (a type of cryptophytes) exhibit buds below the ground, often in the
form of storage organs like bulbs, tubers, or rhizomes. They experience annual
reductions of the entire shoot system.
5. Therophytes are annual plants which die after seed production.
This classification of life-forms based on bud location offers several advantages
from the perspective of fire science. It explains, at least partly, the different strata
within the fuel complex. Moreover, life-forms are also related to post-fire survival
and regeneration strategy. Therophytes will not survive fire, and their persistence is
thus linked to the existence of seed banks. The protection of the buds provided by
belowground storage organs confers geophytes with the capacity to resprout after
the fire. Hemicryptophytes and chamaephytes may, or may not, be able to resprout
after the fire, depending upon the protection of the bud conferred by tissues, and
their regeneration after fire may thus show mixed responses. Resprouting in pha-
nerophytes may be basal or epicormic, depending on bud location and protection,
6.2 Development of Live Wildland Fuels 97

and their post-fire regeneration may originate reproductively from seed or vegeta-
tively, as we will see in Chap. 7.

6.2.2  ariation in Growth and Allocation Patterns Is


V
Coordinated with Post-fire Regeneration Strategy

As fuels grow, different strata develop across the fuel complex, which results partly
from the variety of the co-occurring life-forms. However, ontogenetic development
also regulates fuel structure and its temporal dynamics. The two developmental fac-
tors influencing fuel strata are growth rates and canopy architecture (Table 6.1).
Here we will discuss growth rates, and in the next section, we will cover canopy
architecture. Growth rates differ across life-forms, impact fuel loads, and also deter-
mine the time it takes for a newly emerged fuel (i.e., a newly germinated seedling or
a new resprout) to ascend the fuel layers (ground, surface, and canopy). Relationships
between growth rates and fire are bidirectional since the post-fire reproductive strat-
egy also affects growth rates.
An informal (and preliminary) review of some of the available data in the litera-
ture on the differences in growth rate and its drivers is presented in Fig. 6.2, which
considers only woody plants. Growth is a process that depends on plant size, and
therefore, relative growth rates (RGR) are often reported. This review of the litera-
ture reports a higher RGR in plants that lack the capacity for resprouting (R−), rela-
tive to those with a capacity for resprouting (R+, Fig. 6.2). Differences in RGR
depend on a variety of environmental factors and are tightly linked to the different
attributes (functional traits) of plants. Growth patterns derive from the interaction
between source processes (those that supply carbohydrates, the plant’s building
blocks) and sink activity (the demand for carbohydrates). Ideally, a plant would
only assign resources to build a high leaf area that maximizes photosynthetic carbon
gain. However, the plants also need to develop shoots for light capture and roots to
acquire water and nutrients from belowground resources. Moreover, plants need to
maintain reserves to recover from disturbance and build tissues that survive under
stress. Plants also need to ensure enough resources for reproduction and defense. To
satisfy these competing demands across different plant parts, the internal distribu-
tion of growth (allocation) is such that single resource limitation is minimized
(Chapin et al. 2002).
Source activity, or rates of carbon assimilation, generally correlates poorly with
growth rates because carbon assimilation shows great temporal variation and also
because of control from sink activities. An example of sink control is “compensa-
tory growth,” which refers to the increase in RGR after a perturbation that removed
biomass (e.g., insect infestation that defoliated the tree). Carbon assimilation did
not explain changes in RGR between resprouters and non-resprouters, as differ-
ences were not observed (Fig. 6.2). However, we did observe that nitrogen and
phosphorus concentrations, on a leaf area basis, as well as stomatal conductance,
98 6 Plant Carbon Economies and the Dynamics of Wildland Fuels

)
)

1
1

0.20

25
A (µmol m 2 s
RGR (g g 1 d

15
0.10

5
0.00

R R+ R R+
)
1000

4.0
1

)
g s (mmol m 2 s

3.0
Narea (g m
600

2.0
200

1.0
R R+ R R+
0.10 0.20 0.30 0.40

60
Shoot / Root mass
)2

40
Parea (g m

20
0

R R+ R R+
)3

4.5
Wood density (g cm

Shade tolerance
0.8

3.5
0.6

2.5
1.5
0.4

R R+ R R+

Fig. 6.2 Informal review of the literature highlighting potential differences in relative growth
rates (RGR) between resprouters (R+) and non-resprouters (R−). Such differences derive from the
traits related to resource acquisition and resource use, as well as in allocation patterns. Data from
RGR is sourced from Resco de Dios et al. (2019); data from carbon assimilation (A), specific leaf
area (SLA), nitrogen and phosphorous content on a leaf area basis (Narea, Parea), and stomatal
conductance (gs) are from Maire et al. (2015); shoot-to-root mass ratios are from Falster et al.
(2015); wood density is from Zanne et al. (2009); and shade and drought tolerances are from
Niinemets and Valladares (2006). Significant differences between R− and R+ occurred in RGR,
SLA, Narea, Parea, gs, wood density, shoot/root mass, and shade tolerance

were higher for non-resprouters (Fig. 6.2). These three traits indicate a higher pho-
tosynthetic potential in non-resprouters and less conservative water use, which
could potentially explain higher RGR in these species.
The primary limitations to growth are water and energy (solar radiation).
Belowground growth is higher under water or nutrient limitations, whereas aboveg-
round growth is maximized under light limitation. Non-resprouter plants showed
higher shoot-to-root ratios, a lower degree of shade tolerance and lower wood densi-
ties than resprouters. Lower wood densities allow for faster growth rates as they
6.2 Development of Live Wildland Fuels 99

show lower construction costs. Lower wood densities in non-resprouters are accom-
panied by a higher shoot-to-root ratio because softer woods allow for building
higher plant canopies, as we will discuss in the next section. Higher shoot-to-root
ratios and low wood densities are thus indicative that non-resprouters will be light-­
demanding species (Fig. 6.2).
Other studies suggest that faster growth in non-resprouters could result from
higher hydraulic efficiency (Vilagrosa et al. 2014). Higher efficiency in the water
conducting system may sustain higher rates of stomatal conductance and transpira-
tion that are often required to sustain higher C assimilation rates. Hydraulic effi-
ciency may sometimes come at the cost of decreased drought tolerance because
hydraulic efficiency requires vessels or tracheids, with large diameters, which may
be more vulnerable to cavitation and embolism. However, non-resprouters are often
reported to be more vulnerable to cavitation (Choat et al. 2012) because they have
high shoot-to-root ratios. The shallower roots in non-resprouters limit soil water
availability, which is compensated for by xylem more resistant to cavitation. Non-­
resprouters thus show additional elements in their tissue that allow for greater cavi-
tation resistance, despite high hydraulic efficiency, such as reinforced pit membranes.

6.2.3  lant Development, Canopy Architecture,


P
and Competition as Drivers of Crown Fuel Disposition

Canopy architecture affects the predisposition of canopy fuel particles in space, and
consequently, it is a critical determinant of canopy fires. The amount of biomass in
the crown, its spatial distribution, and the relative proportion of leaves, twigs, and
branches determine the probability of a canopy fire as well as its spread rate. Among
the different parameters affecting fire spread, the canopy bulk density (CBD), which
represents the vertical profile of fuel biomass, is a particularly important driver of
the rate of spread in active crown fires (Table 6.1) (Van Wagner 1977).
The main driver of canopy architecture is the degree of apical dominance, which
indicates the growth rate of the apical meristem in the terminal shoot, relative to
lateral branches. Strong apical dominance is often observed in conifers, which leads
to their typical conical shape. Some hardwoods such as holly (Ilex aquifolium) or
sweet gum (Liquidambar styraciflua) also show strong apical dominance. However,
the rate of growth across the different branches in hardwoods is generally similar,
which leads to rounded crown in those species. Apical dominance may also dimin-
ish with age in some conifer species, such us the Umbrella pine (P. pinea).
Apical dominance is controlled by different growth hormones, which are sub-
stances synthesized in one part of the plant that exert important physiological effects
on a distant part of the plant, even at small concentrations, although the exact mech-
anism is unknown. The degree of apical dominance in conifers is lower in species
from water-limited environments and higher in more radiation-limited environ-
ments (Fig. 6.3). This is probably because higher apical dominance allows for faster
100 6 Plant Carbon Economies and the Dynamics of Wildland Fuels

P. halepensis | P. pinea | P. pinaster | P. nigra | P. sylvestris | P. uncinata

Water limitation

Apical dominance

Fig. 6.3 The degree of apical dominance in Western European pines increases as the degree of
water limitation decreases. Note that P. pinea may have lower apical dominance than P. halepensis
although the latter lives in environments with a higher degree of water limitation. Modified from
J. Luis Ordóñez (2019, CREAF)

vertical growth, which provides a competitive advantage under light limitations.


Consequently, in environments where light is not limiting, plants do not need to
preferentially allocate resources to the apical shoot to avoid competition.
The degree of apical dominance may affect fire behavior, as well as fire-induced
tree damage, because it alters the canopy fuel load and aeration within a canopy: a
given biomass will generally occupy a larger volume under weak apical dominance.
Weaker apical dominance often leads to wider tree crowns, which might also impact
fire behavior (Table 6.1). It is important to remember that the flame front consumes
the thinner fuel particles, often those smaller than 6 mm, while the thicker fuel par-
ticles may be consumed after the flame has passed. Fire spread patterns are deter-
mined by the vertical array of fine fuel particles that are consumed within a fire,
rather than canopy structure as a whole. The array of available canopy fuels has
been proposed to include all foliage plus 50% of the thin (0–6 mm) branch biomass,
but what constitutes available canopy fuel is equivocal (Reinhardt et al. 2006)
because it varies vastly depending on fire intensity and severity.
CBD also results from additional factors such as the pattern of spatial distribu-
tion of leaves within the canopy. Leaves are often distributed aggregated in space
(leaf clumping) or evenly distributed (regular) (Nilson 1971), and random distribu-
tion patterns are rare. The within-canopy pattern of leaf distribution has evolved
jointly to respond to light capture and optimize C assimilation and also to plant
competition (Niinemets and Anten 2009). Species with a higher pattern of aggrega-
tion show a lower light harvesting efficiency but allow for a higher proportion of
6.2 Development of Live Wildland Fuels 101

0.0
Temp Broadleaves

0.1
Macchia

0.2
Clumping index

Trop rain forests


Med Conifers
0.3
0.4

Temp Conifers
Bor Conifers
0.5

eLAI = 2
eLAI = 4.5
0.6

0 2 4 6 8 10
LAI

Fig. 6.4 Distribution of biomes across different combinations of clumping index and leaf area
index (LAI). LAI is directly related to canopy fuel load, and the clumping index is one of the driv-
ers affecting its spatial arrangement. The clumping index varies from 0 to 1. Higher values indicate
a higher degree of clumping and lower values a higher degree of regular distribution. The effective
leaf area index (eLAI) results from the product of LAI and clumping index and indicates the LAI
of a random canopy with equivalent light interception. Data from Cescatti and Niinemets (2004)

foliage in high light, where photosynthetic returns are higher. Leaf clumping scales
positively with leaf area density (Fig. 6.4), which in turn leads to higher CBD, and
species with higher leaf area index show higher degrees of leaf clumping.
Consequently, conifers show a higher degree of clumping and leaf area index than
hardwoods (Niinemets and Anten 2009). The degree of foliage clumping scales
negatively with leaf length, which could exert additional feedbacks on fire severity
as detailed in Sect. 6.3.
Another important factor affecting crown volume is tree diversity and competi-
tion. Some studies have found how the variation in crown volume varies signifi-
cantly with species diversity. Trees occurring at more diverse stands show more
complex canopies than those growing in monocultures (Jucker et al. 2015). This
leads to a more efficient use of the light available, and it also increases the degree of
crown connection. It could thus be argued that tree diversity increases the likelihood
of crown fires. We need further quantifications of how CBD varies in mixed forests,
as most studies have so far been developed in monospecific conifer stands.
Available CBD is currently estimated following a heuristic approach in fire simu-
lators because manual measurements of CBD are very tedious and time-consuming,
leading to an overall lack of data. There have been some attempts towards inferring
CBD from LiDAR (Fernández-Álvarez et al. 2019) or photography (Keane et al.
2005), with promising results, but only CBD values for a few conifer species are
102 6 Plant Carbon Economies and the Dynamics of Wildland Fuels

currently available. Improved estimates of CBD, and data for additional species,
may come from tapping into the physiological or ecosystems ecology literature.
Plant ecophysiologists and ecosystem ecologists are interested in scaling photosyn-
thesis from leaves to canopies, and there is substantial literature on canopy leaf area
profiles for a wider range of species.
Variation in canopy architecture has often been examined as resulting from the
trade-off between light use, competition, and mechanical stability. Canopy architec-
ture is linked with wood density, such that species with denser woods often reach
lower heights, due to increased construction costs and higher mass per unit volume.
Denser wood is also often associated with weaker apical dominance, consistent with
a strategy that maximizes light acquisition in shade-tolerant species. Conversely,
strong apical dominance is often linked with softer wood, where trees grow more
rapidly and reach taller heights in high light environments (Lawton 1984) (Fig. 6.5).
However, a potential role for fire as an agent shaping CBD remains understudied.
There is a very large degree of intraspecific variability in canopy architecture that
subsequently affects CBD. Variability in CBD is strongly affected by competition
and crown class, the hierarchical position of the tree (dominant, co-dominant, inter-
mediate, and suppressed). Given the overarching effect of tree competition on CBD,

Fig. 6.5 Differences in


canopy architecture lead to
differential distribution of
biomass fuels. Species
with weak apical
dominance, denser wood,
lower height, and decurrent
growth habit are often
hardwoods (upper left),
whereas strong apical
dominance, softer wood,
1.0

higher height, and Oak


excurrent growth are often Pine
more prevalent in conifers
0.8

(upper right). Such


differences in canopy
structure may lead to a
Crown length
0.6

differential profile
distribution of foliage
within a canopy (lower
0.4

panel), which then affects


canopy bulk density. Leaf
area density data comes
0.2

from other studies (Jucker


et al. 2015; Kunz et al.
2019)
0.0

0.0 0.2 0.4 0.6 0.8 1.0


Cumulative leaf area biomass
6.3 Seasonal Changes in the Flammability of Live Fuels 103

management plays a key role in determining CBD and reductions in basal area. We
are still lacking a sound theory on how interactions between species identity, abiotic
factors, and competition interact to drive CBD. However, future progress might
come from analyses of large-scale LiDAR analyses.
A crucial discussion these days on wildland fire behavior is whether differences
in species traits are relevant or ecosystem structure is more important. One of the
major assumptions underlying the trait hypothesis is that a given species shows a set
of “fixed” and unchanged functional properties. CBD shows a wide degree of intra-
specific variation, as a result of stand management, which is at odds with the notion
that interspecific variation in functional traits is a relevant driver of landscape fire
patterns.

6.3 Seasonal Changes in the Flammability of Live Fuels

Plant growth and development alter fuel load and structure over annual or decadal
time scales. However, fuel flammability also varies seasonally, and its effects may
be due to changes in water availability, as discussed in Chap. 5, or due to changes in
the carbon economy of the fuels. Changing carbon investments within live fuels
during a growing season may alter stand flammability via two processes: (1) physi-
ological synthesis, storage, and emission of biogenic volatile organic compounds
and (2) leaf and shoot development and aging, which alter inorganic chemistry and
leaf anatomy.

6.3.1  iogenic Volatile Organic Compounds Have


B
an Uncertain Effect over Fire Behavior

Plant species produce a wide range of secondary metabolites, including essential


oils, resins, and biogenic volatile organic compounds (BVOCs) which show very
low ignition values and high heating content, potentially increasing fuel flammabil-
ity. BVOCs represent a varying portion of the plant carbon balance, ranging between
0.1% and 15% of total photosynthates, depending on species and growing condi-
tions (Kesselmeier et al. 2002). Their synthesis and emission rates change through
the year, often peaking during the hottest part of the year, but water scarcity might
limit emissions. BVOCs are emitted by a wide range of species and may also be
stored in resin ducts, cavities, canals, trichomes, or specialized glands in some (e.g.,
Eucalyptus, Pinus, Cistus), but not all (e.g., oaks), species that produce BVOCs
(Laothawornkitkul et al. 2009).
There are different species of BVOCs, but isoprene and terpenes (particularly
monoterpenes) are the most common. Examples of isoprene emitters include
Populus, Salix, Picea, or Eucalyptus, and among the monoterpene emitters, we find
104 6 Plant Carbon Economies and the Dynamics of Wildland Fuels

Pinus, Quercus, or Cistus. BVOCs serve a wide a range of physiological actions,


including, but not limited to, pollinator attraction, deterring pathogens, and herbi-
vores, and protecting the leaves from abiotic stresses such as high temperature or
oxidative stress (Kesselmeier et al. 2002).
A long-standing hypothesis states that BVOCs may enhance flammability by
lowering the energy input required during ignition (Laothawornkitkul et al. 2009).
However, tests for this hypothesis are still scarce, and current evidence on its influ-
ence on live fuel flammability derives from a few studies that assessed the effects of
BVOCs on leaf-level flammability. These studies have observed mixed results, with
some reporting that the effects of BVOCs may be statistically significant but weaker
than those of water content (Alessio et al. 2008; De Lillis et al. 2009), others show-
ing a moderate effect (Della Rocca et al. 2017), and others indicating a predominant
effect of BVOCs concentration over water content (Owens et al. 1998). Unfortunately,
leaf-level processes often scale poorly at the ecosystem or landscape level (as dis-
cussed in Chap. 9), and such tests provide valuable but limited information on the
role of BVOCs as drivers of fire behavior.
Tests on the effects on BVOCs on stand flammability at the level of whole stands
are yet to be developed. Because BVOCs may also be stored in some species, some
studies indicate that stored BVOCs in litter may increase the flammability of the
litter bed (Ormeño et al. 2009).
However, some authors have speculated that BVOC emissions could accelerate
fire spread. BVOCs are denser than air and accumulate near the ground surface.
BVOC emissions increase with temperature, reaching a peak around 150–200 °C. The
increase in BVOC emissions in the pre-heating phase of a fire could lead to a gas
build-up that would lower ignition times and potentially facilitate fire spread
(Chetehouna et al. 2013), and in some particular topographic conditions, it could
also be a contributing factor to eruptive fire behavior (Barboni et al. 2011). It is
important to note that evidence on this hypothesis is circumstantial and that this
hypothesis requires much further testing.

6.3.2  hanges in Fuel Inorganic Chemistry and Leaf Anatomy


C
That Potentially Affect Fire Behavior
6.3.2.1 Inorganic Chemistry

Wildland fuels are primarily composed of cellulose, hemicellulose, lignin, and min-
eral content in different proportions across fuel particles and species (Table 6.2), in
addition to the resins and volatiles discussed in Sect. 6.3.1.
Each of these components exhibits very different heating values, which is the
amount of energy released during combustion (Table 6.2), hence affecting fire
spread. From the lens of fire behavior, the amount of mineral content, or ash, is
particularly important. Ashes are difficult to ignite and dampen fire intensity (Varner
et al. 2015), while fuels with a larger proportion of mineral content tend to burn
6.3 Seasonal Changes in the Flammability of Live Fuels 105

Table 6.2 Chemical composition of different plant material


Lignin (%) Hemicellulose (%) Cellulose (%) Ash (%)
Hardwood stems 18–25 24–40 40–55 <1
Softwood stems 25–35 25–35 45–50 <1
Leaves 0 80–85 15–20 2–8
Grass straw 14–21 24–38 29–43 2–8
Bark 44 30 25 10

slower and have a higher proportion of smoldering (flameless) combustion (Keane


2015). In fact, the mode of action of fire retardants is to diminish fire spread by
increasing the fuel mineral content (Giménez et al. 2004; Keane 2015).
Mineral content is the proportion in dry weight of inorganic or mineral material
(i.e., without C, H, and O). It is often measured as the ratio of the weight of the ash
that is left after a complete combustion, relative to the weight of the unburned dry
fuel particle. The five most abundant components in ash are silicon, potassium,
calcium, sulfur, and chlorine (Sorek et al. 2014). These are essential macronutrients
that play different major physiological roles including defense against pathogens
(silicon), regulation of stomatal behavior (potassium or chlorine), amino acid for-
mation (sulfur), or cell wall construction (calcium), among many others.
The contribution of ash varies markedly across tissues: ~10% in bark (Zaja et al.
2018), 2–8% in leaves (Petisco et al. 2008), and < 1% in wood (Zaja et al. 2018). It
also varies across species, as pine leaves often show lower ash content (2–6%) than
angiosperms (4–8%) (Petisco et al. 2008). The proportion of ash content is not con-
stant in time, and it often shows a unimodal temporal pattern. Ash content increases
rapidly during leaf development due to increased accumulation times. However,
leaf mineral content often diminishes before leaf or branch senescence because
plants have evolved an efficient system of nutrient recycling and minerals stored in
older leaves or branches are remobilized and stored or used in other plant parts prior
to abscission. Some studies have argued that the reduction in ash content during leaf
senescence may have a more profound effect on leaf flammability than the reduction
in leaf moisture (Broido and Nelson 1964), but this hypothesis needs further testing,
and it is not consistent with modeling results (Fig. 6.6).
After leaf and branch senescence and shedding, an increase in the mineral con-
tent often occurs during decomposition in the ground. Mineral content in duff and
litter may be as high as 10–50% (Keane et al. 2012), which leads to a slower rate of
fire spread with a higher proportion of burning in smoldering combustion. The high
concentration of mineral content is thus a contributing factor to explain why surface
fires, particularly those that are propagated through litter and duff, burn at smaller
intensity than crown fires, although other factors such as fuel load and bulk density
also play a role.
Mineral content is a parameter that is often included in surface fire models. For
instance, total mineral content serves to adjust fuel loading. The effective mineral
content, which is defined as the total mineral content minus silica content, is
included as a parameter dampening Rothermel’s reaction intensity (see Chap. 2,
106 6 Plant Carbon Economies and the Dynamics of Wildland Fuels

Fig. 6.6 Ash and live fuel

1.0
Mineral
moisture contents are
Fuel Moisture
incorporated in

0.8
Rothermel’s fire behavior

Damping coefficient
model as damping
coefficients affecting the

0.6
reaction intensity. This
graph compares the effects
of silica-­free mineral

0.4
content and fuel moisture
on the damping coefficient.
The point on the graph

0.2
indicates a commonly used
value in models 0.0

0.00 0.02 0.04 0.06 0.08 0.10


Concentration of silica free mineral content
0.05 0.10 0.15 0.20 0.25 0.30
Fuel Moisture Content

Fig. 6.6). However, this parameter is often static, and its temporal variability is not
included, an assumption that may be valid for dead fuels. It remains untested how-
ever whether seasonal changes in leaf mineral content alter fire behavior in both
surface and canopy fires.

6.3.2.2 Changes in Leaf Anatomy and Organic Chemistry

Seasonal fuel dynamics and subsequent fire behavior may also respond to changes
in leaf anatomical properties (Table 6.1). Among these, leaf thickness and the ratio
of leaf mass per unit area (LMA) is particularly important. LMA varies across func-
tional types, and for instance, sclerophyll trees show higher LMA than other hard-
woods. LMA in deciduous trees often increases at the beginning of the season, due
to the build-up of cell walls, but then it remains stable throughout much of the sea-
son until it drops again prior to senescence (Poorter et al. 2009). In evergreens,
LMA sometimes (but not always) increases with age, due to continued thickening
and lignification of cell walls. LMA is also affected by the concentration of non-
structural carbohydrates, which also fluctuates seasonally. Within canopy patterns
of LMA also vary, and LMA often declines under shade, an adaptation that opti-
mizes photosynthetic carbon uptake. Furthermore, a plethora of environmental fac-
tors affect LMA.
Changes in leaf size and LMA affect fuel behavior because they alter the surface-­
area-­to-volume ratio (SAVR), the area of a particle surface divided by the volume of
that particle. SAVR represents the effect of the fuel size on the combustion process
and the response of fuel particles to variation in temperature and moisture (Keane
2015). SAVR is in turn an important driver of different aspects of fire behavior
6.4 Development of Dead Wildland Fuels 107

(Table 8.1) in Rothermel’s fire behavior model such as fireline intensity or the resi-
dence time (but see Finney et al. (2013)). Changes in leaf size and LMA also alter
the packing ratio of the fuel (the fuelbed bulk density divided by the particle density,
dependent on SAVR) which, in turn, affect the degree of aeration and the rate of
spread according to some (Table 8.1), but not all, models of fire spread (the effect of
SAVR over rate of spread is less important in other modeling approaches). Changes
in leaf size and LMA also alter the leaf fuel moisture content that is expressed on a
dry matter basis, and consequently, increases in LMA, which indicate increases in
dry weight, lead to declines in leaf moisture (please see Chap. 5 for a more compre-
hensive discussion).
Another relevant leaf-level trait altering seasonal fuel dynamic is leaf length,
which increases early in the season and has been linked with fire severity (Schwilk
and Caprio 2011). The effects of leaf length and area seem to be more important
following leaf abscission as a driver of surface fires because shorter or smaller
leaves tend to pack more closely and create a less aerated layer of litter, which hin-
ders propagation.

6.4 Development of Dead Wildland Fuels

6.4.1 Leaf Senescence

The ultimate fate of all living fuels is to become dead fuels. During leaf senescence
there is remobilization of nutrients, which are stored in the stem or roots and are
used for the production of new tissue. Chlorophyll is also degraded in the process,
which results in different leaf colors (Koike 1990). Changes in dead to live fuel
ratios are important drivers of fire behavior (Table 6.1) mostly (but not only as dis-
cussed here) because leaf moisture declines markedly: senesced leaves have a much
lower water content (often below 15%) than green leaves (often >80%). There are
different processes underlying this transformation: aging-related programmed cell
death, drought deciduousness, and drought-induced cavitation. Additionally, plant
death after pathogen or pest attacks also lead to leaf senescence and subsequent
shedding. The potential direct effects of senescence on stand flammability will
depend on whether leaves senesce during or after the fire season.
Aging-related programmed leaf senescence is the last stage of leaf development
due to a complex molecular regulatory network that enhances plant fitness (Lim
et al. 2007). Aging-related senescence follows different temporal patterns across
functional types and species. In winter deciduous species, leaf senescence and shed-
ding occur after the summer fire season and are of scarce relevance for crown fire
behavior. However, leaf dropping enhances litter fuel loads and, consequently,
enhances the probability of a surface fire in subsequent seasons.
Aging-related leaf senescence and subsequent shedding may occur during the
summer in some evergreen species. This is the case, for instance, of Pinus ­halepensis,
108 6 Plant Carbon Economies and the Dynamics of Wildland Fuels

one of the most common pine species in the Western Mediterranean Basin. Pinus
halepensis needles often live up to 3 years, and the shedding of the older leaf cohort
occurs at the onset of the fire season, between May and July, depending on environ-
mental conditions (Fig. 5.3). Dead leaf and floret retention also occur in other spe-
cies such as those from the genus Banksia (Proteaceae) or Xanthorrhoea preissii
(Xanthorrhoeaceae), an arborescent grass tree. Dead leaves have a much lower
moisture content than live leaves (< 15% vs > 80%), and consequently, canopy
flammability is greatly enhanced during this period where a substantial proportion
of the canopy could be under such low fuel moisture. The period of dead leaf reten-
tion varies widely among clades, from less than 2 weeks in pines to a few years in
X. preissii. Shedding of dead leaves leads to reductions in canopy bulk density that,
consequently, decreases the likelihood of crown fires.
Drought deciduousness is a reduction in leaf area in response to water shortage
(Smith et al. 1997; Vilagrosa et al. 2003). Reductions in leaf area allow for reduced
water demand and are a common response in hardwood species, including ever-
green species. Some extreme examples occur in desert species, which may become
completely deciduous. Drought deciduousness may be a rapid response through leaf
shedding, which leads to a short-term (i.e., weeks) increase in crown flammability
during the periods between senescence and shedding. Drought deciduousness may
also lead to longer-term allocation adjustments in the ratio of leaf area to sapwood
area, which tends to reduce overall canopy fuel load.
Drought-induced cavitation is another mechanism leading to leaf senescence and
subsequent shedding (Kolb and Sperry 1999), and sometimes, it is also the mecha-
nism underlying drought deciduousness. Reductions in the capacity for transporting
water through the xylem may develop during drought as a result of the introduction
of embolisms (bubbles of air) that lead to cavitation (i.e., nonfunctional) of vessels
and tracheids. Leaves and fine branches, together with roots, are often the organs
most susceptible to cavitation because it is difficult to repair cavitated tissue and,
consequently, tissues with lower construction costs are the first to suffer cavitation
(Zimmermann 1978).
Plants are modular organisms, and cavitation of these thin elements acts like a
fuse to protect the stem and other tissues with higher construction costs (“hydraulic
fuse” hypothesis). Drought-induced cavitation thus starts in leaves and distal
branches, but if drought continues, it could lead to eventual plant death (Tyree and
Sperry 1988). Drought-induced cavitation transiently increases the amount of dead
fuel in the canopy for a few weeks between senescence and shedding. Its effects
may range from being localized in a branch to entire canopies.
Leaf senescence is also the result of pest and pathogen infestations. An extreme
example is the bark beetle attacks across the Rocky Mountains of North America.
Bark beetle leads to complete defoliation, and during a brief period of time after
plant death, the tree retains its red, dry needles until they eventually shed. This could
lead to temporary increases in the availability of crown fuel (Harvey et al. 2014). It
is worth noting that not all pathogen or insect infestations lead to increases in flam-
mability. Defoliating insects such as the pine processionary moth, for instance,
6.4 Development of Dead Wildland Fuels 109

which affects pine stands across Europe, tend to decrease canopy bulk density
through herbivory and, consequently, tend to diminish fire danger.
Although leaf senescence has the potential to increase torching and crown fire
behavior, there is still no evidence to indicate that this potential is being realized in
the field. This is partly due to the lack of studies on this topic, because this is a
parameter that is often not included in crown fire modeling studies. Furthermore,
the degree of leaf senescence is often not quantified in the field, and therefore
addressing its role in fire behavior is problematic. Some exceptions to this, however,
are studies on bark beetle infestations, which have not led to an increase in burned
area after needle senesce during insect infestation (Harvey et al. 2014; Hart et al.
2015), although it has been proposed that future risk for firestorm increases after the
large accumulation of dead and standing biomass (Stephens et al. 2018).

6.4.2  elf-Pruning, Self-Thinning, Bark Shedding, and Plant


S
Death

One of the crucial parameters for crown fire behavior is the height at the base of the
canopy, a parameter that is affected by the degree of self-pruning. As trees grow and
the lower branches become shaded, the carbon balance of these branches eventually
becomes negative; that is, carbon uptake through photosynthesis is lower than res-
piration and maintenance costs. The response of the tree to shading in the lower
branches is to induce leaf and branch senescence, a process known as self-pruning
that helps to optimize resource uptake and use (Henriksson 2001). Some studies
have shaded single branches and documented branch senescence, but when the
same intensity of shade was applied to the entire plant, plant death was not observed
(Henriksson 2001). This indicates that branch senescence after shading is not the
direct result of the stress but that the plant appears to “compute” a carbon balance
and “decides” to induce senescence in those parts of the plant. Self-pruning is more
precocious and active in shade-intolerant species and also under denser stands.
Fire spread may be rapidly accelerated by spot fires, which result from embers
and firebrands that are lofted by the convective plume and create a new fire ahead of
the fire front (Table 8.1). The effectiveness of firebrands to create a spot fire and
their dispersal depend upon the interactions between the size of the firebrand and
the surface-to-volume ratio of the firebrand. For instance, Ganteaume et al. (2011)
proposed that (1) heavy firebrands such as pine cones may sustain flames for longer
periods of time and are efficient for long-distance spotting; (2) light but thick fire-
brands (with low surface-to-volume ratio) such as twigs or acorns may be efficient
in short-distance and, sometimes, also in long-distance spotting; and (3) light and
thin (high surface-to-volume ratio) firebrands, such as needles or Eucalyptus bark,
combust quickly and are thus efficient for short-distance spotting (leaves and thin
barks). The role of eucalypt bark as a firebrand has received particular attention
because some authors consider that the decorticating bark from eucalypts favors fire
110 6 Plant Carbon Economies and the Dynamics of Wildland Fuels

spread. However, tests for this hypothesis are scarce, and whereas some authors
consider it would only be effective in short-range spotting, others consider its aero-
dynamic properties could lead to wide range spotting (Ellis 2011; Ganteaume et al.
2011). Alternative views indicate that decorticating bark serves as a protection
against epiphytes (Bowman et al. 2014).
As the stands develops, there is also natural plant mortality that results from
competition in crowded stands. This process, termed self-thinning, is important as it
increases ladder fuels and it thus contributes to developing crown fires. Self-thinning
is a function of biomass accumulation, and it results from plant competition. The
relationship between biomass and density is defined by a log-log relationship, with
a coefficient of 3/2 (Westoby 1984). Self-thinning is important because it creates
ladder fuels that may increase the probability of a crown fire.

6.4.3 Deposition and Decomposition

Canopy fuels change strata and become surface fuels after deposition and finally
become ground fuels after decomposition (Fig. 6.1). The ecological literature often
refers to litterfall, rather than deposition, but this term may be confusing from a fire
perspective as litter refers to a specific fuel type (Keane 2015). Leaf, twig, branch,
or log deposition results from the processes mentioned in Sects. 6.4.1 and 6.4.2.
Leaf deposition is a regular process, and often with a homogeneous dispersal pat-
tern, that may be additionally influenced by interactions between particle size and
shape, wind, and micro-topography. Deposition of thicker particles such as logs
often occur near the tree from where it originates and occurs at more irregular inter-
vals than leaf deposition (Keane 2015).
Deposition rates change markedly across ecosystems and fuel types. Across the
western US conifer forests, they may occur at 0.4 kg m−2 year−1, which is about 3%
of total aboveground production (Keane 2015). In Australian ecosystems, total val-
ues for deposition may vary from 0.3–0.4 kg m−2 in dry sclerophyll forests to
1 kg m−2 in rainforests (Thomas et al. 2014). These values are particularly high as
they are close to, and sometimes higher than, 100% of the annual net primary pro-
ductivity. Deposition rates are positively affected by both temperature and
precipitation.
Decomposition rates vary markedly across different ecosystem types, but also
with different fuel characteristics. Decomposition is a three-step process (Keane
2015) that begins with (1) the leaching of soluble organic compounds that are
washed with the water and filter into the soil; followed by (2) the physical break-
down, or fragmentation, of plant material into smaller pieces which is often carried
out by invertebrate soil fauna; and finally by (3) microbial respiration that is respon-
sible for the chemical breakdown of the remaining detritus.
The effects of precipitation and temperature on decomposition are not linear, like
in the case of deposition. Decomposition depends on microbial activity, which is
inhibited under water logging, as well as under high aridity. Whereas fuel p­ roduction
References 111

and subsequent deposition are low in arid environments, fuel production and depo-
sition are high in very wet environments. Such high fuel production and low decom-
position rates may lead to large accumulation of ground fuels, particularly in peat
ecosystems.
The final fate of deposited fuel is to burn, rot, or be eaten. Some studies have
examined potential relationships between traits affecting flammability and those
affecting decomposition. In a sample of 32 evergreen perennials from eastern
Australia, Grootemaat et al. (2015) observed that traits affecting flammability at
bench scale were unrelated to those altering decomposition rates. Lignin contribu-
tion is one of the most important drivers of decomposition rates. However, factors
related to leaf size and shape, which led to different types of packing, were more
related to flammability.
It has, for instance, been shown that, on a bench, the small leaves of junipers and
other non-Pinus Pineacea lead to a surface that is poorly ventilated and difficult to
ignite and carry a flame (although landscape patterns of fire behavior may be very
different from these). Small broadleaved species leaves may lead to fuelbeds too
loosely packed to be flammable. Interestingly, plant diversity may increase flam-
mability, as the merging of small needles and small broadleaved species was more
flammable (Zhao et al. 2019).
Not only do flammability and decomposability at bench scales vary independent
from each other, but also flammability in different plant organs follows different
paths. It has been observed, for instance, different species rankings in terms of flam-
mability depend upon whether the ranking is based on leaves, bark, or shoot (Zhao
et al. 2019).
However, it is important to note that these discussions regarding links between
flammability and decomposability are limited to studies performed under artificial
conditions, such as experimental fuelbeds. In the field, we can expect to observe a
tighter connection between flammability and decomposability such that the latter, at
least to a point, affects the former. That is, species with low decomposition rates will
tend to accumulate higher fuel loads, and consequently, they will lead to more flam-
mable surfaces. Fire models often assume that fuel loads are a major driver of fire
behavior, and other factors related to size, such as particle diameter, surface volume
to area ratio, and others previously mentioned in this chapter, also play a role
(Table 6.1). However, as discussed in more detail in Chap. 9, factors other than spe-
cies physiological traits are often more important in driving extreme fire behavior.

References

Albini FA (1976) Estimating wildfire behavior and effects. USDA Forest Service, Intermountain
Research Station, General technical report INT-30 Ogden, UT, 23 pp
Alessio GA, Peñuelas J, Llusià J, Ogaya R, Estiarte M, Lillis MD (2008) Influence of water
and terpenes on flammability in some dominant Mediterranean species. Int J Wildland Fire
17:274–286
112 6 Plant Carbon Economies and the Dynamics of Wildland Fuels

Barboni T, Cannac M, Leoni E, Chiaramonti N (2011) Emission of biogenic volatile organic com-
pounds involved in eruptive fire: implications for the safety of firefighters. Int J Wildland Fire
20:152–161
Bowman DMJS, French BJ, Prior LD (2014) Have plants evolved to self-immolate? Front Plant
Sci 5:590. https://doi.org/10.3389/fpls.2014.00590
Broido A, Nelson MA (1964) Ash content: its effect on combustion of corn plants. Science
146:652–653
Cescatti A, Niinemets Ü (2004) Leaf to landscape. In: Smith WK, Vogelmann TC, Critchley C
(eds) Photosynthetic adaptation: chloroplast to landscape. Springer, New York
Chapin FS, Matson PA, Mooney HA (2002) Principles of terrestrial ecosystem ecology. Springer,
New York
Chetehouna K, Courty L, Mounaïm-Rousselle C, Halter F, Garo J-P (2013) Combustion char-
acteristics of p-cymene possibly involved in accelerating forest fires. Combust Sci Technol
185(9):1295–1305. https://doi.org/10.1080/00102202.2013.795557
Choat B, Jansen S, Brodribb TJ, Cochard H, Delzon S, Bhaskar R, Bucci SJ, Feild TS, Gleason
SM, Hacke UG, Jacobsen AL, Lens F, Maherali H, Martinez-Vilalta J, Mayr S, Mencuccini M,
Mitchell PJ, Nardini A, Pittermann J, Pratt RB, Sperry JS, Westoby M, Wright IJ, Zanne AE
(2012) Global convergence in the vulnerability of forests to drought. Nature 491(7426):752–
755. https://doi.org/10.1038/nature11688
De Lillis M, Bianco PM, Loreto F (2009) The influence of leaf water content and isoprenoids on
flammability of some Mediterranean woody species. Int J Wildland Fire 18:203–212
Della Rocca G, Madrigal J, Marchi E, Michelozzi M, Moya B, Danti R (2017) Relevance of ter-
penoids on flammability of Mediterranean species: an experimental approach at a low radiant
heat flux. iForest Biogeosci For 10(5):766–775. https://doi.org/10.3832/ifor2327-010
Ellis PFM (2011) Fuelbed ignition potential and bark morphology explain the notoriety of the
eucalypt messmate ‘stringybark’ for intense spotting. Int J Wildland Fire 20:897–907. https://
doi.org/10.1071/WF10052
Falster DS, Duursma RA, Ishihara MI, Barneche DR, FitzJohn RG, Vårhammar A, Aiba M, Ando
M, Anten N, Aspinwall MJ, Baltzer JL, Baraloto C, Battaglia M, Battles JJ, Bond-Lamberty
B, van Breugel M, Camac J, Claveau Y, Coll L, Dannoura M, Delagrange S, Domec J-C,
Fatemi F, Feng W, Gargaglione V, Goto Y, Hagihara A, Hall JS, Hamilton S, Harja D, Hiura T,
Holdaway R, Hutley LS, Ichie T, Jokela EJ, Kantola A, Kelly JWG, Kenzo T, King D, Kloeppel
BD, Kohyama T, Komiyama A, Laclau J-P, Lusk CH, Maguire DA, le Maire G, Mäkelä A,
Markesteijn L, Marshall J, McCulloh K, Miyata I, Mokany K, Mori S, Myster RW, Nagano
M, Naidu SL, Nouvellon Y, O'Grady AP, O'Hara KL, Ohtsuka T, Osada N, Osunkoya OO, Peri
PL, Petritan AM, Poorter L, Portsmuth A, Potvin C, Ransijn J, Reid D, Ribeiro SC, Roberts
SD, Rodríguez R, Saldaña-Acosta A, Santa-Regina I, Sasa K, Selaya NG, Sillett SC, Sterck F,
Takagi K, Tange T, Tanouchi H, Tissue D, Umehara T, Utsugi H, Vadeboncoeur MA, Valladares
F, Vanninen P, Wang JR, Wenk E, Williams R, de Aquino Ximenes F, Yamaba A, Yamada T,
Yamakura T, Yanai RD, York RA (2015) BAAD: a biomass and allometry database for woody
plants. Ecology 96(5):1445–1445. https://doi.org/10.1890/14-1889.1
Fernández-Álvarez M, Armesto J, Picos J (2019) LiDAR-based wildfire prevention in WUI: the
automatic detection, measurement and evaluation of Forest fuels. Forests 10(2):148
Finney MA, Cohen JD, McAllister SS, Jolly WM (2013) On the need for a theory of wildland fire
spread. Int J Wildland Fire 22(1):25–36. https://doi.org/10.1071/wf11117
Ganteaume A, Guijarro M, Jappiot M, Hernando C, Lampin-Maillet C, Pérez-Gorostiaga P,
Vega JA (2011) Laboratory characterization of firebrands involved in spot fires. Ann For Sci
68(3):531–541. https://doi.org/10.1007/s13595-011-0056-4
Giménez A, Pastor E, Zárate L, El P, Arnaldos J (2004) Long-term forest fire retardants: a review
of quality, effectiveness, application and environmental considerations. Int J Wildland Fire
13(1):15
References 113

Grootemaat S, Wright IJ, van Bodegom PM, Cornelissen JHC, Cornwell WK, Schweitzer J (2015)
Burn or rot: leaf traits explain why flammability and decomposability are decoupled across
species. Funct Ecol 29(11):1486–1497. https://doi.org/10.1111/1365-2435.12449
Hart SJ, Schoennagel T, Veblen TT, Chapman TB (2015) Area burned in the western United States
is unaffected by recent mountain pine beetle outbreaks. Proc Natl Acad Sci USA 112:4375–
4380. https://doi.org/10.1073/pnas.1424037112
Harvey BJ, Donato DC, Turner MG (2014) Recent mountain pine beetle outbreaks, wild-
fire severity, and postfire tree regeneration in the US northern rockies. Proc Natl Acad Sci
111(42):15120–15125. https://doi.org/10.1073/pnas.1411346111
Henriksson J (2001) Differential shading of branches or whole trees: survival, growth, and repro-
duction. Oecologia 126(4):482–486. https://doi.org/10.1007/s004420000547
Jucker T, Bouriaud O, Coomes DA, Baltzer J (2015) Crown plasticity enables trees to opti-
mize canopy packing in mixed-species forests. Funct Ecol 29(8):1078–1086. https://doi.
org/10.1111/1365-2435.12428
Keane RE (2015) Wildland fuel fundamentals and applications. Springer, Cham
Keane RE, Reinhardt ED, Scott J, Gray K, Reardon J (2005) Estimating forest canopy bulk density
using six indirect methods. Can J For Res 35:724–739. https://doi.org/10.1139/x04-213
Keane RE, Gray K, Vacciu V (2012) Spatial variability of wildland fuel characteristics in Northern
Rocky Mountain ecosystems. Res Pap RMRS-RP-98. US Department of Agriculture, Forest
Service, Rocky Mountain Research Station, Fort Collins, CO, 56 p
Kesselmeier J, Ciccioli P, Kuhn U, Stefani P, Biesenthal T, Rottenberger S, Wolf A, Vitullo M,
Valentini R, Nobre A, Kabat P, Andreae MO (2002) Volatile organic compound emissions in
relation to plant carbon fixation and the terrestrial carbon budget. Glob Biogeochem Cycles
16(4):73-71–73-79. https://doi.org/10.1029/2001gb001813
Koike T (1990) Autumn coloring, photosynthetic performance and leaf development of deciduous
broadleaved trees in relation to forest succession. Tree Physiol 7:21–32
Kolb KJ, Sperry JS (1999) Differences in drought adaptation between subspecies of sagebrush
(Artemisia tridentata). Ecology 80(7):2373–2384
Kunz M, Fichtner A, Hardtle W, Raumonen P, Bruelheide H, von Oheimb G (2019) Neighbour
species richness and local structural variability modulate aboveground allocation patterns and
crown morphology of individual trees. Ecol Lett. https://doi.org/10.1111/ele.13400
Laothawornkitkul J, Taylor JE, Paul ND, Hewitt CN (2009) Biogenic volatile organic compounds in
the earth system. New Phytol 183(1):27–51. https://doi.org/10.1111/j.1469-8137.2009.02859.x
Lawton RO (1984) Ecological constraints on wood density in a tropical montane rain forest. Am J
Bot 71(2):261–267. https://doi.org/10.1002/j.1537-2197.1984.tb12512.x
Lim PO, Kim HJ, Nam HG (2007) Leaf senescence. Annu Rev Plant Biol 58:115–136. https://doi.
org/10.1146/annurev.arplant.57.032905.105316
Maire V, Wright IJ, Prentice IC, Batjes NH, Bhaskar R, van Bodegom PM, Cornwell WK,
Ellsworth D, Niinemets Ü, Ordonez A, Reich PB, Santiago LS (2015) Global effects of soil
and climate on leaf photosynthetic traits and rates. Glob Ecol Biogeogr 24(6):706–717. https://
doi.org/10.1111/geb.12296
Niinemets Ü, Anten NPR (2009) Packing the photosynthetic machinery: from leaf to canopy. In:
Laisk A, Nedbal L, Govindjee (eds) Photosynthesis in silico understanding complexity from
molecules to ecosystems, Advances in Photosynthesis and Respiration. Springer, Dordrecht,
pp 364–399
Niinemets Ü, Valladares F (2006) Tolerance to shade, drought, and waterlogging of tem-
perate northern hemisphere trees and shrubs. Ecol Monogr 76:521–547. https://doi.
org/10.1890/0012-9615(2006)076[0521:TTSDAW]2.0.CO;2
Nilson T (1971) A theoretical analysis of the frequency of gaps in plant stands. Agric Meteorol
8:25–38
114 6 Plant Carbon Economies and the Dynamics of Wildland Fuels

Ordóñez JL (2019) Cómo diferenciar los pinos autóctonos peninsulares. In Justamante


A (ed) Claves para identificar las diferentes especies de pinos, ¡infografía dis-
ponible [Blog post]. 27th May 2019. Retrieved from http://blog.creaf.cat/es/noticias/
claves-para-identificar-las-diferentes-especies-de-pinos-infografia-disponible/
Ormeño E, Céspedes B, Sánchez IA, Velasco-García A, Moreno JM, Fernandez C, Baldy V (2009)
The relationship between terpenes and flammability of leaf litter. For Ecol Manag 257(2):471–
482. https://doi.org/10.1016/j.foreco.2008.09.019
Owens MK, Lin C-D, Taylor CA, Whisenant SG (1998) Seasonal patterns of plant flammability
and monoterpenoid content in Juniperus ashei. J Chem Ecol 24(12):2115–2129. https://doi.org
/10.1023/A:1020793811615
Petisco C, García-Criado B, Vázquez de Aldana BR, García-Ciudad A, Mediavilla S (2008) Ash and
mineral contents in leaves of woody species: analysis by near-infrared reflectance spectroscopy.
Comm Soil Sci Plant Anal 39(5–6):905–925. https://doi.org/10.1080/00103620701881253
Poorter H, Niinemets Ü, Lourens P, Wright IJ, Villar R (2009) Causes and consequences of varia-
tion in leaf mass per area (LMA): a meta-analysis. New Phytol 182(3):565–588
Raunkiaer C (1934) The life forms of plants and statistical plant geography (H. Gilbert-Carter &
A. Fausbøll, trans. English ed.). Oxford University Press, Oxford
Reinhardt E, Scott J, Gray K, Keane R (2006) Estimating canopy fuel characteristics in five coni-
fer stands in the western United States using tree and stand measurements. Can J For Res
36(11):2803–2814. https://doi.org/10.1139/x06-157
Resco de Dios V, Chowdhury FI, Granda E, Yao Y, Tissue DT (2019) Assessing the potential func-
tions of nocturnal stomatal conductance in C3 and C4 plants. New Phytol 223:1696–1706
Rothermel RC (1972) A mathematical model for predicting fire spread in wildland fuels. USDA
Forest Service, Intermountain Forest and Range Experiment Station, Research paper INT-115.
Ogden, UT
Schwilk DW, Caprio AC (2011) Scaling from leaf traits to fire behaviour: community com-
position predicts fire severity in a temperate forest. J Ecol 99(4):970–980. https://doi.
org/10.1111/j.1365-2745.2011.01828.x
Smith SD, Monson RK, Anderson JE (1997) Drought-deciduous shrubs. In: Physiological ecol-
ogy of North American Desert plants, Adaptations of desert organisms. Springer, Berlin/
Heidelberg, pp 109–123
Sorek N, Yeats TH, Szemenyei H, Youngs H, Somerville CR (2014) The implications of
Lignocellulosic biomass chemical composition for the production of advanced biofuels.
Bioscience 64(3):192–201. https://doi.org/10.1093/biosci/bit037
Stephens SL, Collins BM, Fettig CJ, Finney MA, Hoffman CM, Knapp EE, North MP, Safford
H, Wayman RB (2018) Drought, tree mortality, and wildfire in forests adapted to frequent fire.
Bioscience 68(2):77–88. https://doi.org/10.1093/biosci/bix146
Thomas PB, Watson PJ, Bradstock RA, Penman TD, Price OF (2014) Modelling surface fine fuel
dynamics across climate gradients in eucalypt forests of South-Eastern Australia. Ecography
37(9):827–837. https://doi.org/10.1111/ecog.00445
Tyree MT, Sperry JS (1988) Do woody plants operate near the point of catastrophic xylem dys-
function caused by dynamic water stress? Answers from a model. Plant Physiol 88(3):574–580
Van Wagner CE (1977) Conditions for the start and spread of a crown fire. Can J For Res 7:23–34
Varner JM, Kane JM, Kreye JK, Engber E (2015) The flammability of forest and woodland litter: a
synthesis. Curr For Rep 1(2):91–99. https://doi.org/10.1007/s40725-015-0012-x
Vilagrosa A, Bellot J, Vallejo VR, Gil-Pelegrin E (2003) Cavitation, stomatal conductance, and
leaf dieback in seedlings of two co-occurring Mediterranean shrubs during an intense drought.
J Exp Bot 54(390):2015–2024. https://doi.org/10.1093/jxb/erg221
Vilagrosa A, Hernandez EI, Luis VC, Cochard H, Pausas JG (2014) Physiological differences
explain the co-existence of different regeneration strategies in Mediterranean ecosystems. New
Phytol 201(4):1277–1288. https://doi.org/10.1111/nph.12584
Westoby M (1984) The self-thinning rule. In: MacFadyen A, Ford ED (eds) Advances in ecological
research, vol 14. Academic Press, pp 167–225. https://doi.org/10.1016/S0065-2504(08)60171-3
References 115

Zaja G, Szyszlak-Bargłowicz J, Gołebiowski W, Szczepanik M (2018) Chemical characteristics of


biomass ashes. Energies 11:2885
Zanne AE, Lopez-Gonzalez G, Coomes DA, Ilic J, Jansen S, Lewis SL, Miller RB, Swenson NG,
Wiemann MC, Chave J (2009) Global wood density database. Dryad. Identifier: http://hdl.
handle.net/10255/dryad.235.
Zhao W, Logtestijn RSP, Hal JR, Dong M, Cornelissen JHC (2019) Non-additive effects of leaf
and twig mixtures from different tree species on experimental litter-bed flammability. Plant
Soil 436:311–324. https://doi.org/10.1007/s11104-019-03931-3)contains
Zimmermann MH (1978) Hydraulic architecture of some diffuse-porous trees. Can J Bot
56:2286–2295
Chapter 7
Effects of Fire on Plant Performance

Abstract Fire impacts many aspects of plant performance and functioning. The
first one is survival, which will be mostly driven by the degree of leaf and bud com-
bustion in non-resprouters experiencing aboveground fires. Secondary reasons
underlying fire-induced plant death include the degree of cambium charring, per-
cent losses of hydraulic conductance, or root consumption. Understanding fire sur-
vival in resprouters is more challenging. Traditional hypotheses on an overarching
role of stored reserves as drivers of resprouting vigor have not always been sup-
ported by data. Alternative views on the drivers of post-fire resprouting include the
degree of drought-induced hydraulic limitations prior to the fire as a major driver.
Cambium charring and hydraulic failure may lead to legacy effects affecting subse-
quent plant responses to drought and to other post-fire stresses. Low- to mid-­
intensity wildfires may induce plant defenses for protection against insect attacks,
but pests are a major reason underlying post-fire tree death. Individuals surviving
the fire, particularly in low- or medium-severity fires, may show enhanced growth
rates due to competition removal, which increases water availability, and, poten-
tially, to transient increases in nutrient availability. Such competition removal may
also enhance reproduction in some trees, and fire acts as a cue to induce flowering
in many geophytes and resprouters.

7.1 Introduction

Many plant species succumb during or after the passage of a fire. Here we define a
plant as dead once it has lost its photosynthetic capacity. That is, we consider that a
plant dies at the moment when (i) it is 100% defoliated and (ii) it has lost the capac-
ity to develop new leaves. Strictly speaking, a completely defoliated plant will still
show some metabolic activity, such as maintenance respiration. However, once it
has lost its capacity to generate new foliage, the complete cessation of metabolic
activity is only a matter of time.
Fire-induced plant mortality has been extensively reviewed in the literature
(Karavani et al. 2018; Bar et al. 2019; Michaletz and Johnson 2007; Hood et al.
2018; Midgley et al. 2010). Here we seek to compile these diverging views into a

© Springer Nature Switzerland AG 2020 117


V. Resco de Dios, Plant-Fire Interactions, Managing Forest Ecosystems 36,
https://doi.org/10.1007/978-3-030-41192-3_7
118 7 Effects of Fire on Plant Performance

single framework and expand it towards also addressing fire effects on growth, leg-
acy effects, and reproduction of surviving trees.
Mortality rates differ markedly across species and functional types (endogenous
factors) and also with fire regime attributes (exogenous factors). Post-fire environ-
mental conditions further modulate mortality rates, and stressful conditions, for
instance, may further exacerbate the negative effects of the fire on plant perfor-
mance. The major endogenous factor determining post-fire survival is the capability
for resprouting after fire, although other structural and physiological adaptations
also modulate the response (Table 7.1). The major exogenous factors are related to
fire regime attributes such as the type (e.g., crown vs surface), intensity, and fre-
quency of fire (Table 7.1). In this chapter we will first discuss the mechanisms
explaining plant death in non-resprouters. We will divide the discussion between
direct effects and indirect or lagged responses. Post-fire mortality may occur either
immediately after the fire or up to a few years afterwards. Therefore, we present a
framework that explicitly incorporates direct fire effects along with indirect fire
responses or legacy effects on plant survival (Fig. 7.1). We will then discuss the
physiological and structural adaptations that may help non-resprouting plants sur-
vive a fire. Next, we will discuss the limits to fire-induced resprouting. The chapter
ends with a discussion on indirect fire effects on plant growth and longer-term sur-
vival and reproduction.

Table 7.1 Relationship between fire regime attributes and critical effects leading to tree mortality
Fire regime attributes Critical effect
Propagation Physiological Survival
type Intensity Frequency Severity process affected by
Threshold
Surface Low Frequent Low, but Cambium Bark Bark thickness
depends on and phloem thickness
LT0 > 2 cm
bark necrosis LT50 < 1 cm
LT100 < 0.1 cm
High Infrequent Depends on Leaf and bud Bud/leaf Crown
% scorch size, canopy defoliation
defoliation architecture, LT0 < 70–30%
Crown High Infrequent High, but Leaf and bud FMC LT50 40–90%
depends on combustion LT100 > 80–90%
%
defoliation
Ground Low Infrequent High, but Root Surface-to-­ Cambium
depends on consumption volume root temperature
time-­ ratio, bark LT0 < 40 °C
integrated thickness LT50 ~ 50 °C
temperature LT100 > 70 °C
LT0, LT50, and LT100 refer to the lethal thresholds (LT) that eliminate 0%, 50%, and 100% of the
population, respectively. LT0 thus indicates complete survival
7.2 Direct Effects of Fire on Plant Performance 119

Fig. 7.1 Different degrees


of varying fire intensity 1
affect fire severity which, Varying fire
in turn, impact different intensity
aspects of plant
performance affects

1
Varying fire
severity

0
affects

1 Mortality

Tree 0
performance
-1

7.2 Direct Effects of Fire on Plant Performance

Fires ultimately kill trees because of the (sometimes interacting) effects of combus-
tion, the interruption of downward carbon transport, and the interruption of sap
transport (Fig. 7.2).
Combustion is the primary mechanism underlying tree death in crown and high-­
intensity fires (Table 7.1), and death results from the elimination of aboveground
biomass and, particularly, of leaves and buds. The flow of carbohydrates through the
phloem may be interrupted due to stem charring, which may ultimately induce root
carbon starvation, which is more prevalent in surface fires but also present in crown
fires. The flow of sap may be interrupted due to losses in hydraulic conductance,
which might increase plant susceptibility to future droughts, and it is also more
prevalent in surface fires. Additionally, root consumption may occur in ground fires,
and this also interrupts sap transport.
120 7 Effects of Fire on Plant Performance

Canopy
A
Leaf area

Crown WUE
combustion

Phloem Cambium Xylem Sap


transport
Necrosis
impaired
Necrosis No repair
No repair

NSC transport Cavitation Structural


impaired changes

Root C starvation Mycorrhizae Soil


production
Fine root VWC
production Root area

Fig. 7.2 Direct and indirect fire effects on plant survival. The three main processes underlying tree
death (combustion, starvation, and dehydration) in bold. Fires kill seeder trees directly by crown
combustion, and the plant succumbs once a critical reduction in leaf and bud area is reached. Fire
may also damage phloem (necrosis), cambium (necrosis), and xylem (cavitation) tissues. When
new growth is impaired (no repair), the tree may either die from carbon starvation or dehydration
(sap transport impaired). Continuous lines ending with arrows indicate endogenous factors, and
dotted lines ending with diamonds indicate exogenous factors. Direct fire effects are included in
boxes, but indirect legacy effects may also affect processes in boxes. For instance, a fire might
directly decline fine root area by smoldering combustion. However, a further decline in root area
might result from carbon starvation

7.2.1 Crown Defoliation

The primary and most notorious effect of fire on tree performance is the partial
combustion and elimination of aboveground biomass (Fig. 7.3). In non-resprouting
individuals, the elimination of leaves and buds will inexorably lead to individual
death (van Wagner 1973; Sieg et al. 2006).
High-intensity crown fires burn canopies by convection, leading to widespread
defoliation, and consequently, the survival of non-resprouters is rare. Some high-­
intensity surface fires may generate enough radiation to also scorch entire canopies,
and tree mortality is thus not necessarily restricted to crown fires. There is a large
degree of interspecific variability on the probability of mortality after different
degrees of crown scorch. For instance, Rigolot (2004a) reported, for S France, that
7.2 Direct Effects of Fire on Plant Performance 121

Fig. 7.3 Probability of

1.0
P . h al epensi s from S East France
mortality as a function of P . h al epensi s from NE Spain
% crown volume damaged P . pi nea from S East France
in P. halepensis from

0.8
different locations and P.

Probability of mortality
pinea. Data from Rigolot
(2004b) and Sylla (2017)

0.6
0.4
0.2
0.0

0 20 40 60 80 100
% Crown Volume Damage

Pinus halepensis showed a 50% survival (LT50, lethal threshold for 50% the popula-
tion) when 45% of its canopy had been defoliated, while LT50 occurred at 91%
canopy defoliation in Pinus pinaster at the same site (Fig. 7.3). Substantial intraspe-
cific variation also occurs as Sylla (2017) documented a much higher LT50 of 81%
canopy defoliation in P. halepensis in NE Spain (Fig. 7.3).
It is important to separate the effects of fire on plants depending on their regen-
eration strategy because the effects of crown combustion and phloem charring pre-
vail on non-resprouting trees, whereas hydraulic limitations may be more important
in resprouter trees, as we will discuss later in this chapter (Sect. 7.3.2).
Such dramatic differences in LT50 across and within species are related to differ-
ent structural and physiological attributes. One of the main factors explaining the
highest LT50 of P. pinaster, relative to P. halepensis, is the larger leaf and bud size
which confers a higher thermal insulation and, consequently, a higher crown sur-
vival (Fernandes et al. 2008). Additional factors such as differences in canopy archi-
tecture, which affect crown spread, and leaf humidity, which affects crown spread
and also thermal insulation, may also play an additional role.

7.2.2 Phloem and Cambium Charring

Forest fires may also char the phloem and cambium. This is the result of protein
denaturation, which occurs almost instantaneously at 60 °C or higher or during
prolonged exposures to temperatures over 50 °C. Such high temperatures are due to
122 7 Effects of Fire on Plant Performance

conduction during the fire front or to flameless combustion after the fire front has
passed. The degree of heating is not uniform across the bole. It is higher at the lee-
ward side, where the flame height is higher and the residence time longer, poten-
tially leading to fire scars at the base (as explained in Chap. 9).
The effects of phloem charring are similar to those of stem girdling as the down-
ward flow of carbohydrates is interrupted leading to root C starvation. Cambium
necrosis impairs the regeneration of the phloem as well as the development of new
xylem cells.
Cambium necrosis is often considered as a secondary factor driving post-fire tree
mortality. Bark provides thermal insulation, and it is sometimes assumed that the
effects of fire on stem temperature stay below critical temperatures once a critical
thickness of 2 cm is reached. However, the relationship between bark thickness and
thermal insulation is further modulated by water content, bark density, and the pro-
portion of inner to outer bark (most thermal protection is provided by the outer
bark), such that important differences between species occur (Wesolowski et al.
2014; Gill and Ashton 1968; Brando et al. 2012). For example, the time to reach
60 °C in different trees with a bark of 1 cm exposed to a propane torch varied
between 118 sec in Acer rubrum, 136 sec in Quercus nigra, and 208 sec in Pinus
palustris. Part of this variation could be induced by torch itself, as it does not heat
uniformly. To circumvent this problem, Sparks et al. (2017) propose the use of dose-­
response relationships, where the response to different doses of fire radiative energy
is examined. It is logistically challenging to measure directly thermal insulation in
the field while controlling exactly the radiation that is applied, but experiments of
this type will allow for more detailed characterization of differences in bark insulat-
ing properties.
Bark thickness increases linearly with stem diameter (see Chap. 4). Thus, even if
the main stem is often protected against critical fire temperatures, thinner branches
will exhibit thinner barks, and consequently, important girdling will occur in canopy
fires or in high-intensity surface fires where radiation leads to high canopy tempera-
tures. Branch or stem girdling through cambium necrosis could explain why tree
mortality occurs when as little as 45% of the canopy has been defoliated (Fig. 7.3).
Mortality in species where canopy defoliation is far from 90% often shows a lagged
response, and death may occur 2 or more years after the fire. This is consistent with
a C starvation response, as trees often show enough reserves to survive without new
C inputs for a few years (Karavani et al. 2018).

7.2.3 Fire-Induced Cavitation and Sap Transport Impairment

Heat plume-induced losses in hydraulic conductivity have long been recognized


(Rundel 1973; Ducrey et al. 1996; Balfour and Midgley 2006), but their effect was
traditionally considered as of minor importance relative to cambium necrosis.
Furthermore, it was only in the last decade when formal demonstrations that fires
could induce catastrophic xylem failure occurred (Michaletz et al. 2012; West et al.
7.3 Resprouting 123

2016; Kavanagh et al. 2010; Bar et al. 2018). There are two complementary
­mechanisms that explain how the heat plume may induce cavitation. The first is via
air seeding cavitation after the extreme vapor pressure deficits created by the heat
plume. Branches or stems often become nonfunctional and lose their capacity to
transport the sap once 80% of their hydraulic conductivity is lost (Resco et al. 2009;
Urli et al. 2014; Hammond et al. 2019). After reaching this critical threshold, the
tree is unable to recover its conductive tissue. The second mechanism is the defor-
mation of cell walls in water-conducting vessels or tracheids after heating cells to
temperatures over 60 °C, although evidence for this process is equivocal (West
et al. 2016).
Forest fires thus lead to substantial reductions in sapwood area and sap flux den-
sity when bark is not thick enough to insulate the xylem from critical temperatures.
These reductions in hydraulic conductivity may subsequently lead to desiccation of
other tree parts such as the crown (Tyree and Zimmerman 2002). Plant death from
dehydration is a more rapid process than C starvation, occurring typically within
weeks, and it may consequently explain rapid post-fire mortality.

7.2.4 Root Consumption

Soil surface temperatures may reach 1000 °C or more during a high-intensity fire,
but soils are poor thermal conductors, and often little change in temperature is
recorded below the uppermost 5–15 cm, (Choczynska and Johnson 2009), particu-
larly in the dry summer months. Soil and root heating primarily occur by conduc-
tion during smoldering combustion (Hood et al. 2018).
Potential fire effects on roots have seldom been assessed, and it is currently
assumed that root damage is restricted to ground fires, where thermal stress occurs
during extended periods of time. The primary mechanism underlying ground fire-­
induced death is thus related to root mortality, which ultimately leads to dehydration
and nutrient starvation. Unlike the previously discussed mechanisms, that mostly
affect non-sprouters, root consumption may also prevent resprouting because roots
act as major storage organs (Clarke et al. 2013).

7.3 Resprouting

Resprouting is the most efficient mechanism of resistance to fire in environments


characterized by high-frequency and/or high-intensity crown fires. This is the major
trait conferring fire resistance in angiosperms, but it is rare in gymnosperms.
Resprouting canopies may recover their foliage promptly after the fire, with some
studies indicating recovery times as low as 3–6 years after fire (Caccamo et al.
2015). There is intrinsic variability in the capacity for resprouting within one spe-
124 7 Effects of Fire on Plant Performance

cies. Species classified as resprouters are often those where more than 70% of the
individuals do resprout (Vesk and Westoby 2004).
Clarke et al. (2013) proposed the buds-protection-resources framework to under-
stand resprouting, which is useful to encapsulate the diversity of resprouting types
and responses. According to this framework, resprouting varies across species and
functional types depending on the location of the bud bank, on the bud protection
strategy, and on bud resources.

7.3.1 Location and Protection of the Bud Bank

Bud bank location is often used to differentiate the different resprouting types
which, broadly speaking, may be aerial, basal, or belowground (Fig. 7.4). Aerial
resprouters may be either apical, when the bud is in the terminal meristem, or epi-
cormic, when the buds are in the primary stem or branches (note that additional
types are described in Clarke et al. (2013), and the reader is referred to that publica-
tion for a more complete description). Apical resprouters include arborescent mono-
cots, tree ferns, and cycads, among others (Fig. 7.4). These species protect the
terminal bud by the imbricated network of mature leaf bases and leaf primordia,
which cover and insulate the bud. Strictly speaking, these species should be consid-
ered as apical sprouters, rather than apical resprouters, because they do not resprout
as such.
Epicormic resprouting occurs from dormant buds that are protected by the bark,
and different degrees of protection occur depending on the size of the bud, bark
thickness, and bud distance to the bark surface. Epicormic resprouting is present in

Fig. 7.4 Different types of resprouting based on bud location. (a) Epicormic resprouting in euca-
lypt and apical resprouting in the tree ferns (Photo credit: Rachael H Nolan). (b) Basal resprouting
in different Mediterranean species
7.3 Resprouting 125

some of the predominant genus within the temperate and tropical biomes, such as
Eucalyptus and Quercus.
Basal resprouting may occur from the lignotuber (a swollen root crown), from
the root collar, or from the axillary basal buds in herbaceous species with stolons or
in root bases. Basal resprouters include many epicormic resprouters, but also addi-
tional shrubs from the Mediterranean Basin, and it is also common in savannas,
particularly in Africa. Protection of the buds in basal resprouters is also provided by
the bark or stem bases and soil and, in herbs, by leaf sheaths or clumped stems and
rhizomes.
Belowground resprouting may originate from belowground storage organs such
as bulbs, corms (swollen and subterranean plant stem that grows vertically), rhi-
zomes (modified subterranean stems that grow horizontally), tubers, swollen verti-
cal roots, or root suckers. The bud bank is protected by the soil, which acts as a
buffer against the extreme temperatures occurring during a fire, and additionally by
the bark or leaf bases in geophytes.

7.3.2 Resources Limiting Resprouting

Understanding the drivers and limitations of resprouting has been the topic of much
debate. Resprouting ability depends upon the interactions between a viable bud
bank and bud forming meristematic tissues, the availability of nonstructural carbo-
hydrates (NSC) to support the regrowth until photosynthetic capacity is recovered,
and the availability of water and nutrients (Karavani et al. 2018). Limited bud avail-
ability may limit resprouting in apical sprouters, but basal resprouting is unlikely to
be limited by the bud bank. For instance, Wildy and Pate (2002) documented that a
4-year-old eucalypt seedling already had >1000 buds. Epicormic resprouters are
also unlikely to be limited by the bud bank, except for the rare case where buds get
consumed during the fire.
There is conflicting evidence on the role of carbohydrates as drivers of resprout-
ing (Pate et al. 1990; Bowen and Pate 1993). Plants store large amounts of NSC, and
one of the roles of such a reserves bank could be to provide resources for plant
recovery and new growth after the fire (Wiley and Helliker 2012). NSC could limit
resprouting, for instance, in understory plants that receive little light and conse-
quently have a limited NSC concentration. Additionally, because photosynthetic C
assimilation is impaired under strong drought, NSC reserves could be used to fuel
respiration and deplete the amount of stored reserves. However, it is rare to experi-
mentally observe NSC limitations after drought (Korner 2003; Duan et al. 2014;
Palacio et al. 2014). Eventual carbohydrate depletion could also occur after subse-
quent fire events that trigger resprouting at high frequency (Bowen and Pate 1993)
or under low light (Casals and Rios 2018). However, some studies reported that the
concentrations of nonstructural carbohydrates are not related to resprouting vigor
(Richards and Caldwell 1985; Erdmann et al. 1993; Cruz et al. 2003), and the role
126 7 Effects of Fire on Plant Performance

of storage organs concerning remobilization of other nutrients (e.g., N and P) is also


not clear (Clarke et al. 2013).
Karavani et al. (2018) raised the possibility of observing a “resprout exhaustion
syndrome” after repeated exposure to perturbations or after joint effects of stress
and perturbation. This hypothesis is now starting to be validated by independent
studies (Fairman et al. 2019). The degree of stress experienced prior to the fire could
play an important role in driving the response, especially when important losses in
hydraulic conductivity had developed in the drought preceding the fire. Additional
effects, such as increasing tree defoliation (Carnicer et al. 2011), and processes
declining the health of resprouters, such as the oak decline (Brasier 1996; Gil-­
Pelegrín et al. 2008), could additionally limit resprouting (Fig. 7.5).
This hypothesis is backed up by reports of increasing losses of resilience in
resprouter-dominated ecosystems as stress increases over time (Díaz-Delgado et al.
2002). There are indeed reports of resprouting exhaustion in Mediterranean shrubs
under water (but not carbohydrate) scarcity (Cruz and Moreno 2001), and even
death of resprouters has been documented under acute water stress (Fensham et al.
2009). Moreover, drought acts as a weakening agent that increases the probability
of suffering attacks by pests and pathogens which would substantially hamper tree
resprouting (Oliva et al. 2014). Tree age could act as an additional mechanism that
diminishes resprouting vigor, particularly in unmanaged stands (meaning with older

Fig. 7.5 Declining Quercus ilex (left) and Picea brachytyla (right). Climate change is inducing
forest decline in many areas worldwide. This is often the result of a complex process triggered by
acute water scarcity and where interactions between pests and pathogens along with the impair-
ment of the capacity to transport water and carbon and the consumption of reserves weaken the
tree. Water scarcity enhances fuel availability, and it may also complicate population persistence.
The oak decline observed in this picture is an example of this process, and it is to be expected that
this will be a contributing factor to the resprout exhaustion syndrome. In spruce, it may increase
fire vulnerability
7.4 Indirect Effects of Fire on Plant Performance 127

and larger stems), due to the increasing costs of respiratory maintenance with size
and to hormonal changes (Juvany et al. 2015; Munné-Bosch 2015).

7.4 Indirect Effects of Fire on Plant Performance

7.4.1 Growth and Survival

It is important to remember that fire is not a binary variable (e.g., “burned/unburned”)


and that its effects on tree mortality will depend upon the intensity and the severity
of the burn. Low- to intermediate-intensity fires, either in planned or unplanned
fires, could have positive effects for tree growth and hydraulic status. This is because
low- and intermediate-intensity fires, more common in mountainous regions of the
Mediterranean Basin, for instance (Touchan et al. 2012; Slimani et al. 2014), will
not lead to crown scorch, hydraulic, or cambium damage in mature trees with thick
barks but may reduce competition by killing smaller individuals or the understory
vegetation. It may also increase nutrient availability in the post-fire ash bed. In these
circumstances fire may act as a thinning operation that reduces competition and,
subsequently, temporarily improves the stand’s water status (Resco de Dios 2016;
Alfaro-Sánchez et al. 2016).
Fires may also exert negative carry-over effects due to hydraulic impairment if
the intensity and severity of the fire increase. This possibility, however, is only now
beginning to be examined (Bar et al. 2017). Plume-induced cavitation may lead to
hydraulic losses and also conduit deformation. This would have a negative effect
particularly in those species where xylem remains functional during a few years,
such as pines or diffuse-porous species. Ring-porous species water transport is
dominated by earlywood (wood formed at the beginning of the season). Earlywood
vessels naturally become either embolized or blocked with tyloses at the end of the
growing season (Kitin and Funada 2016). Only latewood remains functional for a
few years, but their role in transporting sap is very minor, because it has a much
smaller diameter and sap flow scales to the power of 4 of the radius. Consequently,
negative carry-over effects due to hydraulics after fire are less likely to occur in spe-
cies with ring-porous xylem.
There might also be negative carry-over effects over growth derived from
decreased whole-tree carbon gain leading to reduced mycorrhizal nutrient intake,
which further declines carbon assimilation. Defoliation reduces carbon assimila-
tion, and consequently, the amount of carbon that is allocated to grow mycorrhiza
could suffer. Reduced C allocation to mycorrhiza could decrease their growth which
would consequently decline nutrient uptake. In fact, Owen et al. (2019) reported
reduced abundance and diversity of ectomycorrhizal fungi even 15 years after a
high-severity fire in a ponderosa pine forest. Such declines in nutrient uptake could
then feedback to further decline C allocation.
128 7 Effects of Fire on Plant Performance

Trees surviving a fire are more susceptible to fungal decay and insect attacks.
Fungal decay further weakens tree defenses and increases susceptibility to insect
attacks. Insect attacks may also act as vectors for fungal propagation, creating a
self-reinforcing feedback ultimately leading to tree death.
Bark beetles are the most common insects that attack surviving trees after a fire.
Bark beetles require that the degree of scorching is high enough to weaken tree
defenses, that the bark is thin enough to allow the attack, but also thick enough to
ensure there is undamaged inner bark for feeding the insects (Parker et al. 2006).
Such insect attacks are also part of the reason why post-fire tree mortality may occur
a few months after the fire. However, low- to mid-intensity fires induce defenses
against insects, and sometimes, they may resist the attack better, as we will discuss
to a greater detail in Chap. 9.

7.4.2 Reproduction

Fire may also serve to stimulate cone production in the years after the fire. A com-
mon response thinning is an increase in cone production, presumably driven by the
removal of competition and the consequent availability of additional resources
(Moreno-Fernández et al. 2013). There is some evidence indicating that a similar
stimulation on cone production also occurs after fire. The evidence on whether or not
fire provides an additional stimulation beyond that observed in response to thinning
in conifers is equivocal (Krannitz and Duralia 2004; Peters and Sala 2008). Some
studies report the opposite pattern, meaning that thinning is not enough to ensure
pine regeneration and that it only occurs when prescribed burning occurred immedi-
ately after the thinning. This was the case, for instance, of Pinus nigra, a non-serot-
inous montane tree adapted to surface fire regimes (Fig. 7.6a) (Tardós et al. 2019).

Fig. 7.6 Fire as a driver of regeneration. (a) A prescribed surface fire promoted seedling establish-
ment in a Pinus nigra forest from NE Spain. (b) Fire-stimulated flowering in the grass tree
Xanthorrhoea at Wilsons Promontory National Park, Australia. (Photo credit: Rachael H Nolan)
References 129

Not only fire presence but also its timing is important. Peters and Sala (2008)
observed how fall burns had a more positive effect on seed viability and mass than
spring burns on Pinus ponderosa. Some species produce serotinous cones, but this
will be discussed in Chap. 8. At any rate, the mechanisms underlying how
­low-­intensity fires affect reproduction are far from being resolved, and further
research is needed.
The effects of fire on seed production of angiosperm trees depend on the tempo-
ral variability of seed production on that species. For instance Funk et al. (2016)
observed that seed production increased markedly 1 year after fire in the masting
species Quercus macrocarpa but that it returned to baseline levels in a species with
highly stable temporal seed production (Q. ellipsoidalis).
Additionally, there are a wealth of resprouting shrub species and geophytes that
show flowering stimulated by fire (Lamont and Downes 2011) and some extreme
examples of species that only produce seed after a fire (Conceição 2018; Groom and
Lamont 2015). In these species, fire serves as a cue to stimulate flowering although
the underlying mechanism remains unresolved (Fig. 7.6b).

References

Alfaro-Sánchez R, Camarero J, Sánchez-Salguero R, Sangüesa-Barreda G, De Las Heras J (2016)


Post-fire Aleppo pine growth, C and N isotope composition depend on site dryness. Trees
Struct Funct 30:581–595. https://doi.org/10.1007/s00468-015-1342-9
Balfour DA, Midgley JJ (2006) Fire induced stem death in an African acacia is not caused by can-
opy scorching. Austral Ecol 31(7):892–896. https://doi.org/10.1111/j.1442-9993.2006.01656.x
Bar A, Nardini A, Mayr S (2017) Post-fire effects in xylem hydraulics of Picea abies, Pinus sylves-
tris and Fagus sylvatica. New Phytol. https://doi.org/10.1111/nph.14916
Bar A, Nardini A, Mayr S (2018) Post-fire effects in xylem hydraulics of Picea abies, Pinus syl-
vestris and Fagus sylvatica. New Phytol 217(4):1484–1493. https://doi.org/10.1111/nph.14916
Bar A, Michaletz ST, Mayr S (2019) Fire effects on tree physiology. New Phytol. https://doi.
org/10.1111/nph.15871
Bowen BJ, Pate JS (1993) The significance of root starch in postfire shoot recovery of the resprouter
Stirlingia latifolia R. Br. (Proteaceae). Ann Bot 72:7–16
Brando PM, Nepstad DC, Balch JK, Bolker B, Christman MC, Coe M, Putz FE (2012) Fire-induced
tree mortality in a neotropical forest: the roles of bark traits, tree size, wood density and fire
behavior. Glob Chang Biol 18:630–641. https://doi.org/10.1111/j.1365-2486.2011.02533.x
Brasier CM (1996) Phytophthora cinnamomi and oak decline in southern Europe. Environmental
constraints including climate change. Ann For Sci 53:347–358
Caccamo G, Bradstock R, Collins L, Penman T, Watson P (2015) Using MODIS data to analyse
post-fire vegetation recovery in Australian eucalypt forests. J Spat Sci 60(2):341–352. https://
doi.org/10.1080/14498596.2015.974227
Carnicer J, Coll M, Ninyerola M, Pons X, Sánchez G, Peñuelas J (2011) Widespread crown condi-
tion decline, food web disruption, and amplified tree mortality with increased climate change-­
type drought. Proc Natl Acad Sci 108:1474–1478. https://doi.org/10.1073/pnas.1010070108
Casals P, Rios AI (2018) Burning intensity and low light availability reduce resprouting ability
and vigor of Buxus sempervirens L. after clearing. Sci Total Environ 627:403–416. https://doi.
org/10.1016/j.scitotenv.2018.01.227
130 7 Effects of Fire on Plant Performance

Choczynska J, Johnson EA (2009) A soil heat and water transfer model to predict below-
ground grass rhizome bud death in a grass fire. J Veg Sci 20:277–287. https://doi.
org/10.1111/j.1654-1103.2009.05757.x
Clarke PJ, Lawes MJ, Midgley JJ, Lamont BB, Ojeda F, Burrows GE, Enright NJ, Knox KJE
(2013) Resprouting as a key functional trait: how buds, protection and resources drive persis-
tence after fire. New Phytol 197:19–35. https://doi.org/10.1111/nph.12001
Conceição AA (2018) A hot case for conservation: Candombá (Vellozia pyrantha), a flammable
plant endemic to a national park is used to make a fire and threatened by fire suppression policy.
J Nat Conserv 45:118–121. https://doi.org/10.1016/j.jnc.2018.08.011
Cruz A, Moreno JM (2001) Seasonal course of total non-structural carbohydrates in the lignotuber-
ous Mediterranean-type shrub Erica australis. Oecologia 128:343–350. https://doi.org/10.1007/
s004420100664
Cruz A, Pérez B, Moreno JM (2003) Plant stored reserves do not drive resprouting of the lignotu-
berous shrub Erica australis. New Phytol 157(2):251–261
Díaz-Delgado R, Lloret F, Pons X, Terradas J (2002) Satellite evidence of decreasing resilience in
Mediterranean plant communities after recurrent wildfire. Ecology 83:2293–2303
Duan H, Duursma RA, Huang G, Smith RA, Choat B, O’Grady AP, Tissue DT (2014) Elevated
[CO2 ] does not ameliorate the negative effects of elevated temperature on drought-induced
mortality in Eucalyptus radiata seedlings. Plant Cell Environ 37(7):1598–1613. https://doi.
org/10.1111/pce.12260
Ducrey M, Duhoux F, Huc R, Rigolot E (1996) The ecophysiological and growth responses of
Aleppo pine (Pinus halepensis) to controlled heating applied to the base of the trunk. Can J
For Res 26:1366–1374
Erdmann TK, Nair PKR, Kang BT (1993) Effects of cutting frequency and cutting height on
reserve carbohydrates in Gliricidia sepium (Jacq.) Walp. For Ecol Manag 57:45–60. https://doi.
org/10.1016/0378-1127(93)90161-F
Fairman TA, Bennett LT, Nitschke CR (2019) Short-interval wildfires increase likelihood of
resprouting failure in fire-tolerant trees. J Environ Manag 231:59–65. https://doi.org/10.1016/j.
jenvman.2018.10.021
Fensham RJ, Fairfax RJ, Ward DP (2009) Drought-induced tree death in savanna. Glob Chang
Biol 15(2):380–387
Fernandes PM, Vega JA, Jiménez E, Rigolot E (2008) Fire resistance of European pines. For Ecol
Manag 256(3):246–255. https://doi.org/10.1016/j.foreco.2008.04.032
Funk KA, Koenig WD, Knops JMH (2016) Fire effects on acorn production are consistent with
the stored resource hypothesis for masting behavior. Can J For Res 46(1):20–24. https://doi.
org/10.1139/cjfr-2015-0227
Gill A, Ashton D (1968) The role of bark type in relative tolerance to fire of three central Victorian
Eucalypts. Aust J Bot 16:491. https://doi.org/10.1071/BT9680491
Gil-Pelegrín E, Peguero-Pina JJ, Camarero JJ, Fernández-Cancio A, Navarro-Cerrillo R (2008)
Drought and forest decline in the Iberian Peninsula: a simple explanation for a complex
Phenomenon? In: Sánchez JM (ed) Droughts: causes, effects and predictions. Nova Science
Publishers, New York, pp 27–68
Groom PK, Lamont BB (2015) Plant life of southwestern Australia: adaptations for survival. De
Gruyter Open, Warsaw
Hammond WM, Yu K, Wilson LA, Will RE, Anderegg WRL, Adams HD (2019) Dead or dying?
Quantifying the point of no return from hydraulic failure in drought-induced tree mortality.
New Phytol. https://doi.org/10.1111/nph.15922
Hood SM, Varner JM, van Mantgem P, Cansler CA (2018) Fire and tree death: understanding and
improving modeling of fire-induced tree mortality. Environ Res Lett 13(11):113004. https://
doi.org/10.1088/1748-9326/aae934
Juvany M, Müller M, Munné-Bosch S (2015) Bud vigor, budburst lipid peroxidation, and hor-
monal changes during bud development in healthy and moribund beech (Fagus sylvatica L.)
trees. Trees 29(6):1781–1790. https://doi.org/10.1007/s00468-015-1259-3
References 131

Karavani A, Boer MM, Baudena M, Colinas C, Díaz-Sierra R, Pemán J, de Luís M, Enríquez-­


de-­Salamanca Á, Resco de Dios V (2018) Fire-induced deforestation in drought-prone
Mediterranean forests: drivers and unknowns from leaves to communities. Ecol Monogr
88:141–169
Kavanagh KL, Dickinson MB, Bova AS (2010) A way forward for fire-caused tree mortality pre-
diction: modeling a physiological consequence of fire. Fire Ecol 6:80–94
Kitin P, Funada R (2016) Earlywood vessels in ring-porous trees become functional for water trans-
port after bud burst and before the maturation of the current-year leaves. IAWA J 57:315–331
Korner C (2003) Carbon limitation in trees. J Ecol 91(1):4–17
Krannitz PG, Duralia TE (2004) Cone and seed production in Pinus ponderosa: a review. West
North Am Natur 64:208–218
Lamont BB, Downes KS (2011) Fire-stimulated flowering among resprouters and geophytes
in Australia and South Africa. Plant Ecol 212(12):2111–2125. https://doi.org/10.1007/
s11258-011-9987-y
Michaletz ST, Johnson EA (2007) How forest fires kill trees: a review of the fundamental biophysi-
cal processes. Scand J For Res 22:500–515. https://doi.org/10.1080/02827580701803544
Michaletz ST, Johnson EA, Tyree MT (2012) Moving beyond the cambium necrosis hypothesis
of post-fire tree mortality: cavitation and deformation of xylem in forest fires. New Phytol
194(1):254–263. https://doi.org/10.1111/j.1469-8137.2011.04021.x
Midgley JJ, Lawes MJ, Chamaillé-Jammes S (2010) Savanna woody plant dynamics: the role of
fire and herbivory, separately and synergistically. Aust J Bot 58:1–11
Moreno-Fernández D, Cañellas I, Calama R, Gordo J, Sánchez-González M (2013) Thinning
increases cone production of stone pine (Pinus pinea L.) stands in the Northern Plateau (Spain).
Ann For Sci 70(8):761–768. https://doi.org/10.1007/s13595-013-0319-3
Munné-Bosch S (2015) Senescence: is it universal or not? Trends Plant Sci 20(11):713–720.
https://doi.org/10.1016/j.tplants.2015.07.009
Oliva J, Stenlid J, Martinez-Vilalta J (2014) The effect of fungal pathogens on the water and car-
bon economy of trees: implications for drought-induced mortality. New Phytol 203:1028–1035
Owen SM, Patterson AM, Gehring CA, Sieg CH, Baggett LS, Fulé PZ (2019) Large, high-severity
burn patches limit fungal recovery 13 years after wildfire in a ponderosa pine forest. Soil Biol
Biochem 139:107616. https://doi.org/10.1016/j.soilbio.2019.107616
Palacio S, Hoch G, Sala A, Körner C, Millard P (2014) Does carbon storage limit tree growth?
New Phytol 201:1096–1100
Parker TJ, Clancy KM, Mathiasen RL (2006) Interactions among fire, insects and pathogens
in coniferous forests of the interior western United States and Canada. Agric For Entomol
8:167–189
Pate JS, Froend RH, Bowen BJ, Hansen A, Kuo J (1990) Seedling growth and storage characteris-
tics of seeder and resprouter species of Mediterranean-type ecosystems of S.W. Australia. Ann
Bot 65:585–601
Peters G, Sala A (2008) Reproductive output of ponderosa pine in response to thinning and pre-
scribed burning in western Montana. Can J For Res 38(4):844–850. https://doi.org/10.1139/
X07-203
Resco de Dios V (2016) When fire acts like an irrigation: competition release after burning
enhances growth. Trees 30:579–580. https://doi.org/10.1007/s00468-016-1382-9, https://doi.
org/10.1007/s00468-015-1342-9
Resco V, Ewers BE, Sun W, Huxman TE, Weltzin JF, Williams DG (2009) Drought-induced
hydraulic limitations constrain leaf gas exchange recovery after precipitation pulses in the C3
woody legume, Prosopis velutina. New Phytol 181:672–682
Richards JH, Caldwell MM (1985) Soluble carbohydrates, concurrent photosynthesis and effi-
ciency in regrowth following defoliation: a field study with Agropyron species. J Appl Ecol
22:907. https://doi.org/10.2307/2403239
Rigolot E (2004a) Predicting post-fire mortality of Pinus halepensis Mill. and Pinus pinea L. Plant
Ecol 171:139–151
132 7 Effects of Fire on Plant Performance

Rigolot E (2004b) Predicting postfire mortality of Pinus halepensis Mill. and Pinus pinea L. Plant
Ecol 171:139–151
Rundel PW (1973) The relationship between basal fire scars and crown damage in Giant Sequoia.
Ecology 54:210–213. https://doi.org/10.2307/1934393
Sieg CH, McMillin JD, Fowler JF, Allen KK, Negron JF, Wadleigh LL, Anhold JA, Gibson KE
(2006) Best predictors for postfire mortality of Ponderosa Pine trees in the Intermountain West.
For Sci 52:718–728
Slimani S, Touchan R, Derridj A, Kherchouche D, Gutiérrez E (2014) Fire history of Atlas cedar
(Cedrus atlantica Manetti) in Mount Chélia, northern Algeria. J Arid Environ 104:116–123.
https://doi.org/10.1016/j.jaridenv.2014.02.008
Sparks AM, Smith AMS, Talhelm AF, Kolden CA, Yedinak KM, Johnson DM (2017) Impacts of
fire radiative flux on mature Pinus ponderosa growth and vulnerability to secondary mortality
agents. Int J Wildland Fire 26(1):95. https://doi.org/10.1071/wf16139
Sylla D (2017) Mortalidad de Pinus halepensis Mill. debido a los incendios forestales ocurridos en
Catalunya. MS Thesis, University of Lleida, Lleida
Tardós P, Lucas-Borja ME, Beltrán M, Onkelinx T, Piqué M (2019) Composite low thinning and
slash burning treatment enhances initial Spanish black pine seedling recruitment. For Ecol
Manag 433:1–12. https://doi.org/10.1016/j.foreco.2018.10.042
Touchan R, Baisan C, Mitsopoulos ID, Dimitrakopoulos AP (2012) Fire history in European black
pine (Pinus nigra Arn.) forests of the Valia Kalda, Pindus Mountains, Greece. Tree-Ring Res
68(1):45–50. https://doi.org/10.3959/2011-12.1
Tyree MT, Zimmerman MH (2002) Xylem structure and the ascent of sap, 2nd edn. Springer,
Verlin
Urli M, Delzon S, Eyermann A, Couallier V, Rl G-V, Zavala MA, Porte AJ (2014) Inferring shifts
in tree species distribution using asymmetric distribution curves: a case study in the Iberian
mountains. J Veg Sci 25:147–159
van Wagner CE (1973) Height of crown scorch in forest fires. Can J For Res 3:373–378
Vesk PA, Westoby M (2004) Sprouting ability across diverse disturbances and vegetation types
worldwide. J Ecol 92:310–320
Wesolowski A, Adams MA, Pfautsch S (2014) Insulation capacity of three bark types of tem-
perate Eucalyptus species. For Ecol Manag 313:224–232. https://doi.org/10.1016/j.
foreco.2013.11.015
West AG, Nel JA, Bond WJ, Midgley JJ (2016) Experimental evidence for heat plume-induced
cavitation and xylem deformation as a mechanism of rapid post-fire tree mortality. New Phytol
211(3):828–838. https://doi.org/10.1111/nph.13979
Wildy DT, Pate JS (2002) Quantifying above- and below-ground growth responses of the western
Australian oil mallee, Eucalyptus kochii subsp. plenissima, to contrasting decapitation regimes.
Ann Bot 90:185–197
Wiley E, Helliker B (2012) A re-evaluation of carbon storage in trees lends greater support for
carbon limitation to growth. New Phytol 195:285–289
Chapter 8
Forest Succession, Alternative States,
and Fire-Vegetation Feedbacks

Abstract Wildfires trigger changes in vegetation dynamics if the existing commu-


nity does not resist the fire and succumbs. However, the same vegetation state may
still occur if populations show post-fire resilience, that is, if they are able to regener-
ate after fire. Such regeneration will be a function of interactions between propagule
availability (aerial or soil seed banks or other seed sources nearby), its establish-
ment success (dependent on soil resources, dormancy break, herbivores, and micro-
site conditions), fire recurrence, and mycorrhizal networks. When pre-fire species
fail to survive and regenerate after the fire, changes in the vegetation state may lead
to either a forested or a deforested state, depending on legacies from the previously
established community, climatic conditions and water balance, and other processes.
Climate change may compromise post-fire state changes when newly established
communities cannot survive under the novel environmental conditions expected for
the end of this century. Here we review overall post-fire vegetation changes, with a
particular emphasis on Mediterranean environments.

8.1 Post-fire Changes in Community Dynamics

Changes in post-fire community composition depend on a three-step process where


the fire regime interacts with individual-, population-, and community-level pro-
cesses (Fig. 8.1). The first step necessary for a change in community composition is
that the plants present at the site succumb after the fire. A change in community
dynamics cannot occur when the severity of the fire is not high enough to induce
substantial plant mortality. After plant death, the second step towards community
changes depends on the degree of population resilience. Resilience at the population
level is, in turn, determined by the ability of that population to regenerate after a fire.
Fire recurrence and recruitment strategies play a major influence in this second step
since high fire recurrences may limit recruitment. If individuals die after fire and the
population fails to recover, then a species replacement or state transition is likely.
Such transition depends on burned patch size and the distance to new propagule
sources and the environmental conditions immediately after the fire, among others.

© Springer Nature Switzerland AG 2020 133


V. Resco de Dios, Plant-Fire Interactions, Managing Forest Ecosystems 36,
https://doi.org/10.1007/978-3-030-41192-3_8
134 8 Forest Succession, Alternative States, and Fire-Vegetation Feedbacks

Population
succumbs

Individuals
perish

Fire
Forest state 2

Forest state 1

Deforested state

Fig. 8.1 Conceptual framework for the mechanisms of fire-induced deforestation operating across
three hierarchical scales: resistance in individuals, resilience in populations, and transition to a new
community. The first requirement for fire-induced deforestation is the mortality of individuals.
After individuals perish, the second step towards deforestation requires a limited resilience from
the population, that is, an inability of that species to regenerate after fire. If individuals die after fire
and the population fails to recover, then a species replacement or transition to a new state must
occur. In this third step, the community could switch towards a more advanced successional stage
or towards deforestation depending upon different factors. Reproduced with permission from
Karavani et al. (2018)

When changes in community composition after fire do occur, such change may
lead towards either a more or a less advanced successionary stage. For instance, a
forest dominated by a non-resprouting tree species under fire may see no change in
its structure by the passage of a low-severity fire. This was considered to be the case
of the natural fire regimes in many conifer forests globally, where surface fires
served to maintain the ecosystem and also to foster regeneration of the dominant
species (Sect. 7.4.2, Fig. 7.6). This is often the case also after low-severity pre-
scribed burns that seek to diminish ladder fuels. However, a non-resprouting forest
that burns under a high-severity fire, which kills a large portion of the established
trees, may either transition towards another forest that is more advanced in the suc-
cession dynamics. This forest can also transition into a shrubland or grassland, a
degradation process that we will refer to as deforestation in this chapter.
The transition that is most worrisome is deforestation, because it exerts important
feedbacks in the Earth system and it could create a positive feedback in climate
change. In Mediterranean ecosystems, post-fire deforestation has, so far, been mostly
limited to pine-dominated communities (Torres et al. 2016; Martín-Alcón et al. 2015).
Mediterranean pines lack the capacity to resprout, and their regeneration relies upon
post-fire seed germination. Deforestation in the region could thus be considered as
dependent on the concurrence of three conditions: (i) fatal damage to the trees and to
the seeds (e.g., no aerial or soil seed banks) after the fire; (ii) lack of adult conspecifics
that may act as seed sources within dispersal distances (usually <100 m in pines,
8.2 Resilience to Fire at the Population Level 135

Rodrigo et al. 2007); and (iii) absence of resprouting trees in the understory. When
only (i) and (ii) are met, but oaks are present in the understory, then a pine-to-oak (or
mixed forest) transition is likely. However, when (i) and (ii) concur, and no resprout-
ing trees are present in the understory, deforestation is the final result of this process.
Additionally, and although the bulk of post-fire deforestation so far has occurred
in pine forests, we cannot discard deforestation in resprouting forests also under a
warmer and drier future. This is because, for instance, oak decline (Chap. 7, Fig. 7.5)
is already occurring over a large proportion of the Mediterranean (Gómez-Aparicio
et al. 2017; Peguero-Pina et al. 2015; Barbeta and Peñuelas 2016) and resprouting
tree species are being displaced by seeder and resprouting shrub species under
drought (Peñuelas et al. 2017). Consequently, fire may further promote forest-to-­
shrub state transitions also in oak or resprouter forests.
Such vegetation changes have been classically framed as advancing (pine-to-oak
or pine-to-mixed forest) or reversing (pine-to-shrubland) successional dynamics
(Hanes 1971; Trabaud 1984). This approach presents succession as a linear process
where vegetation advances, in the absence of disturbance, towards an endpoint (cli-
max) and, when disturbances occur, vegetation returns to a previous state along that
linear pathway. However, we prefer to describe them within a state transition frame-
work because it is difficult to predict what a successional trajectory should be at a
given site. According to the traditional model of successional dynamics, based upon
a phytosociological perspective, Quercus would often be the predominant natural
potential vegetation for much of the Mediterranean Basin (Rivas-Martínez and co-­
authors 2005). However, palynological evidence has in many occasions documented
the lack of support for such predictions from phytosociological hypotheses on what
the late-successional vegetation for a given area would be (Carrión and Fernández
2009). Additionally, field studies report nonlinear trajectories, meaning that interac-
tions with local idiosyncrasies may lead to different post-fire communities at a given
site. For instance, Baeza et al. (2007) documented how, for different fire recur-
rences, different shrub communities after fire could arise at Pinus halepensis stands
as a function of local processes such as soils.
Given all these nonlinear responses, involving a high degree of complexity
(Fig. 8.2, Table 8.1), predicting large-scale tree mortality and resprouting capacity
after fire, especially under a climate with increasing aridity, remains an important
challenge (Midgley et al. 2010; Clarke et al. 2013). The first step necessary for a
change in community composition, that is, failure to survive after the forest, has
been previously described (Chap. 7), and here we will only succinctly describe the
main knowns and unknowns associated with this process (Table 8.1). The rest of the
chapter will focus on the remaining two steps (regeneration and state changes).

8.2 Resilience to Fire at the Population Level

Tree mortality occurs when individuals succumb to fire. A high-intensity fire is


more likely to reduce survival rates, thus resulting in widespread mortality. In the
case of stand-replacing fires in non-sprouting forests, population resilience depends
136 8 Forest Succession, Alternative States, and Fire-Vegetation Feedbacks

(a) Size/shape burnt patch


On-site seed banks
Off-site
seed banks
Propagule availability

Fire recurrence Soil resources


Dormancy break

Herbivory/pests
Population resilience
Mycorrhizae
Propagule establishment

State shift
(b)
Understory composition
Resprouter forest
Climatic conditions
Soil resources
Seeder forest

Non-forest

Fig. 8.2 Post-fire vegetation dynamics are highly complex. A multitude of processes interact at
different scales to affect individual resistance (described in Chap. 7), population resilience, and
state transitions. State transitions initiate after a fire kills pre-existing individuals, and a series of
factors affect individual resistance to fire. After individuals perish, stand-level resilience (where
pre-fire species are able to regenerate) depends on fire history and recurrence, propagule availabil-
ity, and establishment (a). If pre-fire individuals perish and populations cannot recover, a state shift
is effectuated. The state will transition to a different forest state or to a deforested state depending
on legacies from pre-fire understory composition, soil resources, and climatic pressures (b).
Straight and dotted lines indicate processes that are well-documented or that have been hypothe-
sized but remain in discussion, respectively. Further discussions, and knowns and unknowns asso-
ciated within each step in text and in Table 8.1. Reproduced from Karavani et al. (2018) with
permission

on the ability to recover through the recruitment of new individuals (Fig. 8.2,
Table 8.1). Successful recruitment of individuals is reliant upon the amount of post-­
fire viable seeds available and subsequent seed germination and on seedling estab-
lishment and survival. Mycorrhizal networks, when present, will favor seedling
establishment.

8.2.1 Propagule Availability

The availability and quantity of viable seeds in a burned area are the major require-
ments for successful population recovery, and the source for recruitment can origi-
nate from on-site or off-site seed banks (Resco de Dios et al. 2007; Martín-Alcón
8.2 Resilience to Fire at the Population Level 137

Table 8.1 Processes affecting state transitions and community change after fire in a forest stand
Process Knowns Unknowns
Individual resistance
Live fuel Remains stable and generally available Modeling remains elusive as well
moisture in evergreen trees but shows seasonal as links with the closely related
fluctuations in shrubs, affecting fuel field of plant water relations. The
availability and, consequently, potentialphenology of leaf senescence could
fire behavior lead to trade-offs between drought
and fire resistance
Crown base Determines probability and spread of a Measuring CBD is notoriously
height/canopy crown fire. It could be a niche difficult, and data available only
bulk density construction mechanism covers a few species. Heuristic
predictions are used elsewhere,
which limits fire behavior
simulations
Plant traits Changes in surface leaf area or Whether and how leaf-level
chemical leaf composition could alter flammability scales to the canopy
leaf flammability level remains contentious
Bark Increasing thickness, density, Understanding how fire protection
smoothness, and water content and interacts with additional bark
decreasing fissures enhance cambial functions, such as water and
protection and subsequent survival in photosynthate storage, as drivers of
surface fires bark ecology and evolution
Roots Direct effects of fire on roots are rare, Main unknowns relate to its role as
given the thermal buffering capabilities a storage organ for resprouting
of soils
Physiological Foliage and bud scorch are the main Whether catastrophic xylem failure
processes: factors driving post-fire mortality in is involved in post-fire death,
scorch, girdling crown fires. Cambium necrosis plays a particularly in low-intensity fires,
and cavitation secondary role or in the individuals dying in the
weeks-months after the fire is under
discussion. Trade-offs between
embolism and fire resistances could
occur
Resprouting Driven by the interaction between the Understanding whether water stress
availabilities of water, nutrients, and antecedent to the fire will
reserves. There are indications of eventually lead to a resprout
reduced resprouting vigor after exhaustion syndrome and what is
successive fire events the most limiting mechanism:
cavitation fatigue vs reserve
exhaustion
Population resilience
Propagule Post-fire regeneration in seeder species Factors driving seed production and
availability relies on seed banks or on seeds from dispersal remain difficult to
adjacent unburned stands. Critical incorporate within mechanistic
distances for pine species have been modeling frameworks
reported to be <100 m. The size of the
burned patch is thus critical in species
with no seed banks
(continued)
138 8 Forest Succession, Alternative States, and Fire-Vegetation Feedbacks

Table 8.1 (continued)


Process Knowns Unknowns
Propagule Interactions between soil resources, Competing hypotheses on pre- vs
establishment dormancy break, herbivores and pests, post-fire conditions make different
and microsite conditions are the drivers predictions on the drivers of
of establishment establishment. Neutral processes
remain understudied in the area
Fire recurrence A single crown fire may induce Lack of resprouting under high fire
deforestation in forests dominated by recurrence has not yet been broadly
seeder trees that lack seed banks. documented, but its potential for
Species with seed banks will need at occurrence is high given
least 8–10 years to become mature and documented declines in resprouting
set seed, and increasing fire recurrences vigor
may subsequently induce deforestation
Mycorrhizal Prompt vegetation establishment after An overall lack of data prevents an
networks fire is additionally required to preserve effective synthesis, but there are
ectomycorrhizal networks, as most tree hints that low-intensity fires, such
species in Mediterranean environments as prescribed burning, could be as
are obligate ectomycorrhizal negative as in large fires
State shifts
Legacies form The presence or absence of resprouting The resprout exhaustion syndrome
pre-fire trees in the understory will be the most may reinforce deforestation even in
composition important requirement for determining stands previously dominated by
whether post-fire state change will be resprouters
to a forest state (dominated by
resprouters) or to a deforested state
Climatic Long-term precipitation patterns Whether resprouters will be able to
conditions and influence the site productivity and thrive under future climatic
water balance consequently the general capability for conditions, of increased stress and
the species to resprout or to produce beyond climatic niches in some
and establish propagules. Short-term cases, deserves further attention.
patterns of precipitation in the next wet Even if they can survive, a reduced
season after a fire will affect seed ability to resprout due to the
establishment resprouting exhaustion syndrome
might lead to an alternative
deforested state that is stable within
ecological timeframes, although the
stability and feasibility of these
states are not totally known
Flammability Different species show different traits, Whether species-level flammability
feedbacks such as standing dead biomass, and attributes scale up to affect fire
those may impact stand-level behavior across landscapes remains
flammability contentious
(continued)
8.2 Resilience to Fire at the Population Level 139

Table 8.1 (continued)


Process Knowns Unknowns
Soil resources Soil nutrients affect seed production Extreme soil depletion could lead
and resprouting vigor. Fires increase to an alternative deforested state
erosion, and while temporally that is stable within ecological
increasing nutrient availability (by timeframes, but a lack of literature
increasing mineralization), they prevents assessments on the
generally decrease total nutrient stability of the transition
concentrations. This effect may be
exacerbated under high fire recurrence.
Moreover, soil nutritional status
diminishes under post-fire shrub
communities, relative to pre-fire forest
communities, further hindering
establishment of less frugal tree species
Interactions across these processes may be viewed in Fig. 8.2. Reproduced with permission from
Karavani et al. (2018)

et al. 2015). Propagule availability is thus affected by different aspects of the fire
regime, including fire intensity and severity (for on-site seed banks) or burned size
patch and shape (for off-site seed banks).
On-site recruitment includes soil or aerial seed banks that have escaped heat
damage due to the insulation properties of soil or of fruits and cones, respectively
(Noble and Slatyer 1980; Lloret 1998). Serotiny, the delayed opening of cones or
fruits in response to heat, is a rare feature among EU Mediterranean species, and
within trees, it is present in only two pine species, i.e., P. halepensis and some popu-
lations of P. pinaster (Tapias et al. 2004; Gil et al. 2009) although, strictly speaking,
these are semi-serotinous species as cones also open regularly in the absence of fire.
Fire directly enhances the opening of serotinous cones resulting in a simultaneous
release of large numbers of seeds, under conditions of no competition and high
radiation (Keeley and Fotheringham 2000). Forest management, by regulating tree
density, plays an additional role in determining seed availability as strong competi-
tion inhibits seed production. Consequently, thinning may enhance seed release (as
is often done in regeneration techniques including shelterwood systems
(Matthews 1991)).
One of the main constraints on propagule availability is seed predation. In some
extreme cases, >80% of all seeds in different stands of Pinus halepensis were pre-
dated within the first 6 months after the fire, ultimately leading to regeneration fail-
ure at that site (Pausas et al. 2004). Seed predators are difficult to model as they
comprise a wide variety of species (including birds, rodents, and ants) and vary
strongly with time and space (Ordóñez and Retana 2004). Consequently, seed pre-
dation acts as another major factor controlling propagule availability that can limit
local recruitment and have cascading effects post-fire plant distribution.
140 8 Forest Succession, Alternative States, and Fire-Vegetation Feedbacks

For Mediterranean seeder tree species lacking seed banks, seed originating
from surviving trees in the marginal boundaries of the fire or in unburned
patches within the fire perimeter may be the only alternative seed source. The
spread of propagules and recolonization of the burned area are distance-
restricted by the seed dispersal ability and by the form, sizes, extent, and spa-
tial configuration of burned and unburned patches (Martín-Alcón and Coll
2016; Ordóñez et al. 2005; Vilà-Cabrera et al. 2011). Isolated patches of mature
trees exhibit enhanced seed production due to reduced competition (Ordoñez
et al. 2006). However, dispersal is limited in pines to distances up to 100 m,
indicating that, for species other than the few that are serotinous, propagule
numbers for post-fire regeneration will generally be limiting in large fires. In
addition to seed source proximity, regeneration of these also depends on the
size of the unburned patch (sometimes called fire refugia) where smaller but
evenly distributed unburned patches are more effective than large unburned
patches (Coop et al. 2019).
Propagule availability is, at least to some degree, decoupled from the intensity of
drought immediately after the fire. This is because seed banks do not depend solely
on current year propagules. On the contrary, seeds accumulate for a number of years
prior to the fire, and consequently, water availability on the year of the fire would
not have a major effect on seed availability. Only if extreme water scarcity limits
seed production for a number of years, a significant impact on propagule availability
could be expected (Fig. 8.3).

Fig. 8.3 (a) Serotinous cones in Banksia open after the fire (Photo credit: Rachael H Nolan). (b)
Post-fire seedling regeneration in a Mediterranean forest from aerial (Pinus halepensis) and soil
(Rosmarinus officinalis) seed banks
8.2 Resilience to Fire at the Population Level 141

8.2.2 Propagule Establishment

Successful propagule establishment requires the concurrence of different conditions


and processes, including adequate moisture and nutrient supply, breakage of dor-
mancy in some instances, microclimatic conditions favorable to the development of
that species, no herbivory, and establishing symbiotic relationships with mycorrhi-
zal fungi (in most Mediterranean woody species). Additionally, establishment
requires fire return intervals that are long enough to allow seedlings to become
established and to become mature individuals, so new seeds may be available when
the following fire occurs. Predicting actual establishment rates therefore remains a
complex challenge, particularly if we are also trying to predict the spatial arrange-
ment of the upcoming population. Requirements for species may change markedly
from germination to survival and keep developing through ontogeny (Resco de Dios
et al. 2014; Quero et al. 2008). Moreover, topographic and climatic variation at
landscape scale and, additionally heterogeneity on resource distribution and her-
bivory at fine, microsite, scales add additional complexity to the process.
Here we will distill all of this complexity by focusing only on three intercon-
nected hypotheses that have gained support over the last years to explain post-fire
establishment (once propagules are available). We choose to focus on these three
hypotheses as they indicate a hierarchy of mechanisms, with different processes
becoming important at different periods. We only discuss hypotheses specific to
post-fire establishment, as additional processes generally affecting establishment
(such as herbivory or pests and pathogens, to name a few) have been addressed
elsewhere (Crawley 1996; Wilkinson and Sherratt 2016).
The first hypothesis is that post-fire plant establishment depends on the existence
of “windows of opportunity” (Quintana et al. 2004; García et al. 2016). Germination
and survival do not occur at the same rate at any given time of the year but vary
seasonally. Mediterranean summers tend to be dry and are commonly followed by
significant rainfall in September or October, which provides a window of opportu-
nity for new plants to recruit and develop an extensive root system before the onset
of cold winter temperatures. There is a multiplicity of factors that could affect the
frequency or duration of such windows of opportunity, including intra-annual cli-
matic variation, amount of litter (from scorched leaves which will fall after the fire),
topographic and edaphic factors, and herbivore densities, to name a few (Pausas
et al. 2003; García et al. 2016; Trabaud et al. 1985; Madrigal et al. 2005).
Windows of opportunity occur temporally, but also spatially, into what are known
as “safe sites,” which are microsites which could offer protection for seedlings, such
as undercanopy spaces that provide higher winter temperatures and protect against
winter frosts, for example (Lamont et al. 1993).
The second hypothesis is the “early-to-rise” hypothesis, which predicts that
seedling survival is primarily driven by the establishment date, such that seedlings
emerging early within a window of opportunity have a higher probability of success
than those emerging at a later stage. The underlying rationale is that the first seeds
to appear have more time to develop and face the upcoming unfavorable seasons
142 8 Forest Succession, Alternative States, and Fire-Vegetation Feedbacks

(cold winters or dry summers) with a bigger root system that allows for more exten-
sive soil exploration and longer and thicker hypocotyl or stems that protect against
damaging temperatures (Woods et al. 2014). This hypothesis contrasts with the
safe-site hypothesis, in that it postulates that the spatial arrangement of seedlings
does not ultimately drive recruitment but that the date of emergence within a win-
dow of opportunity plays a crucial role (de Luis et al. 2008).
Related to the windows of opportunity hypothesis is the “event-dependent”
hypothesis, which states that post-fire plant recruitment is driven by the environ-
mental conditions immediately post-fire (Bond and van Wilgen 1996; Moreno et al.
2011). If, for instance, an intense drought (too intense to allow for seedling estab-
lishment) follows a wildfire, the combination of events may create a legacy such
that the plant cover may remain very sparse for a longer time. Therefore, the event-­
dependent hypothesis states that the conditions at the time of the fire create a long-­
term legacy for the posterior dynamics of the community.
The existence of such legacies may be appreciated in Fig. 8.4 where, up to
8 years after the fire, the composition of the community was dependent upon estab-
lishment success during the first year. Indeed, Fig. 8.4 suggests that, while safe sites
may exist, their influence at the plot scale would be limited and exert small impacts
on the subsequent 8 years after the fire. This is because seedling density in different
years after the fire was linearly dependent on the number of germinates in the first
year. The question then becomes as to what drove initial plant establishment and
initial plant density.
Different initial plant establishments have been observed at a given site, depend-
ing upon when the fire occurred, indicating a major effect of the environmental
conditions immediately post-fire. For instance, Moreno et al. (2011) report that
stand composition 3 years after a fire could be explained by the precipitation in the
season immediately after the fire. Therefore, differences in the environmental con-
ditions of the first year after a fire would create differences in initial establishment
(variation in x-axis of Fig. 8.4); this initial composition creates a legacy, and sur-
vival in the upcoming years is simply a fraction of that in the initial years (resulting
from herbivory, diseases, stress, or other processes). Competition and other density-­
dependent processes are additional important drivers of community dynamics, but
their influence often starts at a later stage, after seedling establishment.

8.2.3 Propagule Survival

Interactions between species physiology, herbivory, fire regime, climate, and site
characteristics will determine the survival of the seedling population. Drought is
often the major determinant of propagule survival, but its effects are mediated by
species physiological traits. Protracted periods of water scarcity lead to catastrophic
xylem failure and also to fatal depletions in carbohydrate reserves (carbon starva-
tion) (McDowell et al. 2008). Additionally, the downward flow of photosynthates
through the phloem may be impaired, particularly if high soil temperatures lead to
8.2 Resilience to Fire at the Population Level 143

)
10

10
2

2
R 2 = 0.93 R 2 = 0.84
Abundance Year 1 (m

Abundance Year 2 (m
8

8
6

6
4

4
2

2
0

0
0 2 4 6 8 10 0 2 4 6 8 10
Emergence Year 1 (m 2) Emergence Year 1 (m 2)
)

)
10

10
2

2
R 2 = 0.97 R 2 = 0.93
Abundance Year 3 (m

Abundance Year 8 (m
8

8
6

6
4

4
2

2
0

0 2 4 6 8 10 0 2 4 6 8 10
Emergence Year 1 (m 2) Emergence Year 1 (m 2)

Fig. 8.4 Strong legacy effects of initial emergence on community dynamics. Data represents aver-
age (± SE) values for six different shrub species in a Mediterranean shrubland after a fire. The
dotted line is the 1:1 relationship, and the straight line is the result of least squares fitting.
Emergence in the first year after the fire explained a significant portion of the variation in the abun-
dance of the different species during the first, second, third, and eighth, respectively, year after a
fire. Original data from de Luis et al. (2008). Reproduced with permission from Karavani et al.
(2018)

stem girdling (Helgerson 1990). These three different processes should not be
viewed as mutually exclusive but as interdependent (McDowell 2011). For instance,
stomatal closure prevents runaway cavitation but also inhibits carbon assimilation.
As drought advances, cuticular or residual water losses remain and continually con-
sume the water stored in the stem capacitors until catastrophic thresholds are crossed
(Resco et al. 2009; Skelton et al. 2017; Blackman et al. 2016). Concomitantly, res-
piration also continuously decreases carbohydrate levels during drought, and
whereas carbon starvation is rare in adult plants, lack of reserves has been docu-
mented in seedlings (Pratt et al. 2014). Seedlings cannot recover from a drought
after crossing a certain threshold of xylem hydraulic failure and also after depleting
carbohydrate reserves to the point where demand exceeds supply (Pratt et al. 2014).
Difficulties in measuring phloem flow prevent from generalizations on its over-
all role.
144 8 Forest Succession, Alternative States, and Fire-Vegetation Feedbacks

Survival in adult plants is often related to stomatal behavior such that anisohy-
dric plants, with loose stomatal control, are considered more likely to suffer from
catastrophic cavitation, whereas isohydric plants, with stronger stomatal control,
are considered to be more prone to carbon starvation (McDowell et al. 2008).
However, such simplifying scheme is likely to have limited predictive power of
survival in post-fire regeneration: apart from stomatal behavior, additional traits
affect the water and carbon balances of the plant, including root/shoot ratio (higher
ratios potentially increasing water uptake), cuticular conductance and stem capaci-
tance (as previously discussed), and drought deciduousness, where complete wilt-
ing would not be linked to mortality as the plant would be able to resprout. Post-fire
seedlings will often be more vulnerable to drought than pre-fire adults because of
lower root/shoot ratios, smaller carbohydrate reserves, and also their xylem possi-
bly being more vulnerable to cavitation (lignification in some species does not occur
in their first years).
Herbivory raises an additional challenge for seedling survival because consump-
tion of plant parts may jeopardize seedling survival (Sagra et al. 2017). Seedlings
show a limited size, storage, and capacity for withstanding herbivory, particularly
under dry conditions that already pre-weaken the seedlings (Moreno and Oechel
1991). The effects of herbivory on seedling survival depend upon some aspects of
the fire regime. This is because high-intensity fires often reduce the number of
resprouts and also delay the time of resprouting. In turn, resprouting later in the
season has been linked with increased herbivory because plants that resprout asyn-
chronously from the main flush (later in the season) are more readily attacked by
herbivores (Moreno and Oechel 1991).
However, and contrary to conventional wisdom, post-fire environments may
sometimes provide conditions that are more favorable to plant establishment. This
is because fire diminishes plant competition and it could, consequently, increase
water availability. Parra and Moreno (2017) observed higher predawn water poten-
tials (indicating higher availability of water in the rhizosphere) and growth rates in
burned (seeder and resprouter) shrubs that had been exposed to a 45% reduction of
rainfall, relative to unburned plants exposed to ambient rainfall. This observation is
intriguing and deserves further exploration, as it would suggest that the bottleneck
for post-fire survival may lie in seedling germination and establishment, rather than
in posterior survival. Additionally, this observation could imply that high-intensity
fire may be a useful strategy to regenerate shrublands or forests, but this hypothesis
is still tentative, and additional observations would be required. However, the effects
of post-fire drought on plant regeneration are also contingent upon the fire regime
and, in particular, upon fire recurrence and intensity.
Different attributes of the fire regime are essential drivers of propagule survival.
In particular, short fire intervals hinder plant regeneration because they do not allow
for seed development and also because regular fire application leads to decreased
resprouting vigor. Consequently, the effect of fire recurrence depends upon the
regeneration mode and species. Oak seedlings are often able to resprout even at
1 year old, although consecutive fires at such early ages are likely to also exhaust
resprouting capacity, where such threshold has not yet been established.
8.2 Resilience to Fire at the Population Level 145

­Semi-­serotinous pine species such as Pinus halepensis require intervals of 5 or


more years to produce viable seeds, and consequently, shorter fire return intervals
will compromise viability. Fire at the seedling or sapling stages in non-serotinous
pine species will likely lead to their local extinction because they will not regenerate
and because their thin barks will not effectively provide cambium even from low-
intensity surface fires.

8.2.4 Mycorrhizal Networks

Most woody species in Mediterranean forests are obligate ectomycorrhizal, mean-


ing that they are not able to complete their life cycle without establishing a symbi-
otic relationship with ectomycorrhizal fungi (Smith and Read 2008). This makes the
remaining mycorrhizal community a critical component of the legacy of a burned
stand (Buscardo et al. 2011). A well-developed, healthy, and diverse ectomycorrhi-
zal network may assist propagules to germinate and grow, as the network of fungal
hyphae will increase their access to water and nutrients (Simard et al. 2013).
Whether or not such a functional ectomycorrhizal network is present will be a
function of the time elapsed since woody vegetation was present (Baar et al. 1999).
This is because mycorrhizae depend upon plants as a carbon source and their pro-
tracted absence may lead to eventual depletion of the ectomycorrhizal community.
For instance, the ectomycorrhizal community could be depleted if colonizing annual
herbaceous species become the prevailing vegetation form after a fire, because these
species normally do not form ectomycorrhizae (Torrecillas et al. 2012). If the woody
plant population takes too long to reestablish at the site, they might encounter a
belowground “desert,” devoid of ectomycorrhizal fungi hyphae. Propagules from
tree species would then germinate and grow, but at much reduced rates. Consequently,
trees may not be competitive, in comparison with the colonizing annuals (which
rely on arbuscular mycorrhizae), and such competition may complicate tree estab-
lishment (Richard et al. 2009). In other words, the longer the time for tree establish-
ment, the higher the difficulty for such establishment because the ectomycorrhizal
community may suffer (Lankau et al. 2015). Presence or lack of a functional ecto-
mycorrhizal community may thus be a stabilizing mechanism either for a prompt
tree establishment or for the establishment of a community of ruderal annual plants,
respectively. Moreover, even low-intensity prescribed burning could have negative
long-lasting impacts on the ectomycorrhizal community (Taudière et al. 2017).
However, currently only a limited number of studies have assessed fire effects on
mycorrhizae, hindering synthesis at the time of writing, but available evidence indi-
cates a negative effect of elevated fire recurrence (Taudière et al. 2017; Tomao
et al. 2017).
146 8 Forest Succession, Alternative States, and Fire-Vegetation Feedbacks

8.3  tate Transitions After Fire and Changes in Community


S
Dominance

8.3.1 Short-Term Transitions

Forests in the Mediterranean region of the EU are dominated by either pines or


hardwoods (typically Quercus) in pure or mixed stands. Pines are obligate seeder
species, but hardwoods are facultative resprouters. This individual-level trait makes
hardwood forests more resilient to state shifts after fire, while pine-dominated forest
stands are more prone to shift to either oak (or mixed) forests or to shrublands
(Fig. 8.1, Table 8.1) (Torres et al. 2016; Martín-Alcón and Coll 2016; Baeza
et al. 2007).

8.3.1.1 Transition to Either Mixed or Oak Forest

The vulnerability of pine stands to state transitions depends on some specific species
traits and site features. A single-fire event has been observed to lead to a state shift
in forests dominated by the montane P. nigra and P. sylvestris (Baeza et al. 2007;
Retana et al. 2002; Martín-Alcón and Coll 2016) and others (Torres et al. 2016).
P. nigra and P. sylvestris show late flowering and lack of serotinous cones (Tapias
et al. 2004), and their seeds do not resist high temperatures (Habrouk et al. 1999;
Lanner 1998). Thus, although these pine species might resist low-intensity surface
fires (Fulé et al. 2008), they succumb under high-intensity defoliating fires (Tapias
et al. 2004; Fernandes et al. 2008; Morales-Molino et al. 2017). Their regeneration
is thus linked to the size and the distance from the unburned patch (Coop et al. 2019),
and it can be particularly problematic under high fire recurrence (Coop et al. 2016).
Other pine species such as P. halepensis and P. pinea (Vallejo and Alloza 1998)
are so-called fire-embracer (sensu Pausas 2015) and usually recover after a single-­
fire event (although there are exceptions (Baeza et al. 2007). P. halepensis is usually
able to recruit after a fire from the seeds stored in its semi-serotinous cones, and
P. pinea shows high resistance to fire (Fernandes et al. 2008), along with seeds pro-
tected by thick cones and coats that can endure high temperatures (Escudero et al.
1999). However, state transitions are very likely when a second fire occurs before
newly recruited pines have reached maturity (7–10 years for P. halepensis) (Thanos
and Daskalakou 2000).
Fire can enhance the establishment of oak species, provided oaks are present at
sufficient densities in the understory before the fire, as oak recruitment is generally
very low due to dispersal limitations. Torres et al. (2016) found that post-fire tree
composition and dominance were highly predictable from pre-fire understory struc-
ture. They studied post-fire forest shifts in stands previously dominated either by
non-serotinous populations of P. pinaster or by the resprouter Q. pyrenaica and
documented the shift from Pinus-dominated to Quercus-dominated forest stands
when pre-fire densities of Q. pyrenaica in the understory exceeded 200 stems ha−1.
8.3 State Transitions After Fire and Changes in Community Dominance 147

Similarly, pine-to-mixed forest transitions occurring in the first 15 years after a


high-intensity crown fire have been documented by Martín-Alcón and Coll (2016).
In this case, stands previously dominated by P. nigra have become a mixed forest
where P. nigra co-dominates the stand with Quercus faginea and Quercus cerrioi-
des. P. nigra was able to establish at the site from the seed provided by mother trees
in unburned patches, and the oaks were present in the understory before the fire.

8.3.1.2 Transition to Shrubland

However, when understory oaks are absent, post-fire change to a non-forested state
likely occurs after trees have succumbed to fire and no seed sources are available
(Baeza et al. 2007; Pausas et al. 2008) or in sites of poor quality (e.g., rocky, south-­
facing, steep slope) (Martín-Alcón and Coll 2016).
The non-forested state may lead to lower soil fertility and higher erodibility
(Karhu et al. 2015; Vieira et al. 2015). Losses in soil fertility after a fire may act as
an additional feedback stabilizing state transitions towards a non-forest state. Forest
fires generally increase the availability (mineralization) of N and P, although the
availability of other nutrients, such as K may decrease (Hinojosa et al. 2016).
However, such change in nutrient availability may be transient because total con-
centrations decline and because burned soils are much more erosion prone, which
fosters further nutrient losses (Vieira et al. 2015; Pausas et al. 2008). Moreover,
even if the quantity of soil organic matter may increase after a fire, its quality dimin-
ishes substantially, especially under high fire recurrence. Once a forest-to-shrubland
transition has occurred, the loss in soil organic matter quality and in fertility creates
a stabilizing feedback for the non-forest state (Mayor et al. 2016). This soil degrada-
tion is further enhanced by drought, as enzyme activity decreases with water scar-
city and consequently decreases mineralization rates (Hinojosa et al. 2016).

8.3.2 Long-Term Dynamics

After a state transition has occurred, the crux will lie in understanding whether the
post-fire ecosystem will remain “locked” in its new state. Shrubland and forests
could both be “stable states” (i.e., not transient within ecologically relevant time
scales) for a given set of environmental conditions (Batllori et al. 2015, 2017). There
are experimental indications that post-fire succession in shrublands remains delayed
or stopped, a process termed arrested succession (Baeza et al. 2006; Acácio and
Holmgren 2014; Acácio et al. 2009; Santana et al. 2010). However, verifying this
hypothesis is challenging, as the time scale of available observations is relatively
short with comparison to the longest-lived late-successional species.
The presence of a positive feedback is necessary to explain the existence of alter-
native stable states, following basic dynamical system theory (DeAngelis et al.
1986). The feedback between ecosystem type and soil fertility, which we previously
148 8 Forest Succession, Alternative States, and Fire-Vegetation Feedbacks

described (Mayor et al. 2016), must thus be accompanied by a feedback between


fire and vegetation composition, analogously to what has been vastly argued for
tropical savanna/forest systems (Hoffmann et al. 2012). At a given site, we could
encounter a “pyrophytic” vegetation state, which promotes recurrent fires and
requires them for its maintenance, or a “pyrophobic” vegetation state, which is less
flammable but does not persist in the presence of fires (Bond and Midgley 2012).
For the Mediterranean Basin, some authors consider that there is some evidence
(but not conclusive) for the existence of such fire-vegetation feedbacks for shru-
bland and pine forests. Areas that have suffered fire in the past are more likely to
experience another fire in the future (Loepfe et al. 2010). Shrublands are often con-
sidered as pyrophytic, as shrubs are highly flammable, thanks to their low water
content, high heat of combustion, fine fuel, and high amounts of standing dead
biomass (Baeza et al. 2006, 2011; De Luis et al. 2004; Pausas and Moreira 2012).
The most common shrub species (e.g., Rosmarinus officinalis, Ulex spp., Cistus
spp., Erica spp.) will either accumulate large, fire-resistant, and long-lasting seed-
banks or show a capacity for vegetative propagation, which allows them to regener-
ate and spread very rapidly after a fire (Santana et al. 2012). Conversely, pine forests
are less flammable, but they are prone to disappear after consecutive fires, as we
described above. This connection between plant traits and regrowth is similar to
what is observed in other Mediterranean ecosystems (Odion et al. 2010) and could
potentially lead to a feedback between floristic composition and fire occurrence.
However, forests including resprouter hardwoods are less easily included in this
framework, as they cannot be described as pyrophobic because of their high fire
resistance. At any rate, potential links between species traits and fire behavior are
far from resolved, and different studies indicate that large wildfire activity is inde-
pendent of land cover type (Fernandes et al. 2016; Nunes et al. 2005; Barros and
Pereira 2014). As we will develop further in Chap. 9, fuel properties are only one of
the many factors affecting wildfire behavior, and they are not always the main one.
Climate change raises an additional challenge to the long-term stability of state
changes. Although some oak forests are now becoming established in previously
pine-dominated stands, some of these species will be out of their climatic niche by
the turn of the century. For instance, Torres et al. (2016) documented that Quercus
pyrenaica is currently replacing Pinus pinaster after a stand-replacing fire, as previ-
ously noted. However, these authors also note a decrease in future habitat suitability
due to climate change, and a virtual local extinction for Quercus pyrenaica is mod-
eled using a species distribution model (Ruiz-Labourdette et al. 2012). Moreover,
these authors also note a minimal climate space reduction for the pine species.
Consequently, replacement of pine stands by oak forests may only represent a tran-
sient stage, and climate change may exert an additional force affecting long-term
state changes.
References 149

References

Acácio V, Holmgren M (2014) Pathways for resilience in Mediterranean cork oak land use sys-
tems. Ann For Sci 71(1):5–13. https://doi.org/10.1007/s13595-012-0197-0
Acácio V, Holmgren M, Rego F, Moreira F, Mohren GMJ (2009) Are drought and wildfires turn-
ing Mediterranean cork oak forests into persistent shrublands? Agrofor Syst 76(2):389–400.
https://doi.org/10.1007/s10457-008-9165-y
Baar J, Horton TR, Kretzer A, Bruns TD (1999) Mycorrhizal colonization of Pinus muricata from
resistant propagules after a stand-replacing wildfire. New Phytol 2:409–418
Baeza MJ, Raventos J, Escarre A, Vallejo VR (2006) Fire risk and vegetation structural dynamics in
Mediterranean shrubland. Plant Ecol 187:189–201. https://doi.org/10.1007/s11258-005-3448-4
Baeza MJ, Valdecantos A, Alloza JA, Vallejo RV (2007) Human disturbance and environmental
factors as drivers of long-term post-fire regeneration patterns in Mediterranean forests. J Veg
Sci 18:243–252. https://doi.org/10.1658/1100-9233(2007)18[243:HDAEFA]2.0.CO;2
Baeza MJ, Santana VM, Pausas JG, Vallejo VR (2011) Successional trends in stand-
ing dead biomass in Mediterranean basin species. J Veg Sci 22(3):467–474. https://doi.
org/10.1111/j.1654-1103.2011.01262.x
Barbeta A, Peñuelas J (2016) Sequence of plant responses to droughts of different timescales: les-
sons from holm oak (Quercus ilex) forests. Plant Ecol Divers:1–18. https://doi.org/10.1080/1
7550874.2016.1212288
Barros AM, Pereira JM (2014) Wildfire selectivity for land cover type: does size matter? PLoS One
9(1):e84760. https://doi.org/10.1371/journal.pone.0084760
Batllori E, Ackerly DD, Moritz MA (2015) A minimal model of fire-vegetation feedbacks and
disturbance stochasticity generates alternative stable states in grassland–shrubland–woodland
systems. Environ Res Lett 10(3):034018. https://doi.org/10.1088/1748-9326/10/3/034018
Batllori E, De Cáceres M, Brotons L, Ackerly DD, Moritz MA, Lloret F (2017) Cumulative effects
of fire and drought in Mediterranean ecosystems. Ecosphere 8:e01906. https://doi.org/10.1002/
ecs2.1906
Blackman CJ, Pfautsch S, Choat B, Delzon S, Gleason SM, Duursma RA (2016) Toward an index
of desiccation time to tree mortality under drought. Plant Cell Environ 39:2342–2345. https://
doi.org/10.1111/pce.12758
Bond WJ, Midgley JJ (2012) Fire and the angiosperm revolutions. Int J Plant Sci 173:569–583.
https://doi.org/10.1086/665819
Bond WJ, van Wilgen BW (1996) Fire and plants. Chapman and Hall, London
Buscardo E, Freitas H, Pereira JS, De Angelis P (2011) Common environmental factors explain both
ectomycorrhizal species diversity and pine regeneration variability in a post-fire Mediterranean
forest. Mycorrhiza 21(6):549–558. https://doi.org/10.1007/s00572-011-0363-5
Carrión JS, Fernández S (2009) The survival of the ‘natural potential vegeta-
tion’ concept (or the power of tradition). J Biogeogr 36:2202–2203. https://doi.
org/10.1111/j.1365-2699.2009.02209.x
Clarke PJ, Lawes MJ, Midgley JJ, Lamont BB, Ojeda F, Burrows GE, Enright NJ, Knox KJE
(2013) Resprouting as a key functional trait: how buds, protection and resources drive persis-
tence after fire. New Phytol 197:19–35. https://doi.org/10.1111/nph.12001
Coop JD, Parks SA, McClernan SR, Holsinger LM (2016) Influences of prior wildfires on vegeta-
tion response to subsequent fire in a reburned Southwestern landscape. Ecol Appl 26(2):346–
354. https://doi.org/10.1890/15-0775
Coop JD, DeLory TJ, Downing WM, Haire SL, Krawchuk MA, Miller C, Parisien M-A, Walker
RB (2019) Contributions of fire refugia to resilient ponderosa pine and dry mixed-conifer for-
est landscapes. Ecosphere 10(7):e02809. https://doi.org/10.1002/ecs2.2809
Crawley MJ (1996) Plant ecology, 2nd edn. Blackwell Publishing. https://doi.
org/10.1002/9781444313642.ch13
150 8 Forest Succession, Alternative States, and Fire-Vegetation Feedbacks

De Luis M, Baeza MJ, Raventós J, González-Hidalgo JC (2004) Fuel characteristics and fire
behaviour in mature Mediterranean gorse shrublands. Int J Wildland Fire 13(1):79–87
de Luis M, Verdu M, Raventós J (2008) Early to rise makes a plant healthy, wealthy, and wise.
Ecology 89:3061–3071
DeAngelis D, Post WM, Travis CC (1986) Positive feedback in natural systems. Springer, Berlin
Escudero A, Sanz MV, Pita JM, Pérez-García F (1999) Probability of germination after heat treat-
ment of native Spanish pines. Ann For Sci 56:511–520
Fernandes PM, Vega JA, Jiménez E, Rigolot E (2008) Fire resistance of European pines. For Ecol
Manag 256(3):246–255. https://doi.org/10.1016/j.foreco.2008.04.032
Fernandes PM, Monteiro-Henriques T, Guiomar N, Loureiro C, Barros AMG (2016) Bottom-up
variables govern large-fire size in Portugal. Ecosystems 19:1362–1375. https://doi.org/10.1007/
s10021-016-0010-2
Fulé PZ, Ribas M, Gutiérrez E, Vallejo R, Kaye MW (2008) Forest structure and fire history in an
old Pinus nigra forest, eastern Spain. For Ecol Manag 255:1234–1242
García Y, Castellanos MC, Pausas JG (2016) Fires can benefit plants by disrupting antagonistic
interactions. Oecologia 182(4):1165–1173. https://doi.org/10.1007/s00442-016-3733-z
Gil L, López R, García-Mateos Á, González-Doncel I (2009) Seed provenance and fire-related
reproductive traits of Pinus pinaster in Central Spain. Int J Wildland Fire 18(8):1003–1009.
https://doi.org/10.1071/WF08101
Gómez-Aparicio L, Domínguez-Begines J, Kardol P, Ávila JM, Ibáñez B, García LV (2017) Plant-­
soil feedbacks in declining forests: implications for species coexistence. Ecology 98(7):1908–
1921. https://doi.org/10.1002/ecy.1864
Habrouk A, Retana J, Espelta J (1999) Role of heat tolerance and cone protection of seeds in the
response of three pine species to wildfires. Plant Ecol 145:91–99
Hanes TL (1971) Succession after fire in the chaparral of Southern California. Ecol Monogr
41(1):27–52. https://doi.org/10.2307/1942434
Helgerson OT (1990) Heat damage in tree seedlings and its prevention. New For 3:333–358
Hinojosa MB, Parra A, Laudicina VA, Moreno JM (2016) Post-fire soil functionality and micro-
bial community structure in a Mediterranean shrubland subjected to experimental drought. Sci
Total Environ 573:1178–1189. https://doi.org/10.1016/j.scitotenv.2016.03.117
Hoffmann WA, Geiger EL, Gotsch SG, Rossatto DR, Silva LCR, Lau OL, Haridasan M, Franco
AC (2012) Ecological thresholds at the savanna-forest boundary: how plant traits, resources
and fire govern the distribution of tropical biomes. Ecol Lett 15(7):759–768. https://doi.
org/10.1111/j.1461-0248.2012.01789.x
Karavani A, Boer MM, Baudena M, Colinas C, Díaz-Sierra R, Pemán J, de Luís M, Enríquez-­
de-­Salamanca Á, Resco de Dios V (2018) Fire-induced deforestation in drought-prone
Mediterranean forests: drivers and unknowns from leaves to communities. Ecol Monogr
88:141–169
Karhu K, Dannenmann M, Kitzler B, Díaz-Pinés E, Tejedor J, Ramírez DA, Parra A, Resco de Dios
V, Moreno JM, Rubio A, Guimaraes-Povoas L, Zechmeister-Boltenstern S, Butterbach-Bahl
K, Ambus P (2015) Fire increases the risk of higher soil N2O emissions from Mediterranean
Macchia ecosystems. Soil Biol Biochem 82:44–51
Keeley JE, Fotheringham CJ (2000) Role of fire in regeneration from seed. In: Seeds
the ecology of regeneration in plant communities, vol 2, pp 311–330. https://doi.
org/10.1079/9780851994321.0311
Lamont BB, Witkowski ETF, Enright NJ (1993) Post-fire litter microsites: safe for seeds, unsafe
for seedlings. Ecology 74:501–512
Lankau RA, Zhu K, Ordóñez A (2015) Mycorrhizal strategies of tree species correlate with trailing
range edge responses to current and past climate change. Ecology 96:1451–1458
Lanner RM (1998) Seed dispersal in Pinus. In: Richardson DM (ed) Ecology and biogeography of
Pinus. Cambridge University Press, Cambridge
Lloret F (1998) Fire, canopy cover and seedling dynamics in Mediterranean shrubland of
Northeastern Spain. J Veg Sci 9:417–430. https://doi.org/10.2307/3237106
References 151

Loepfe L, Martinez-Vilalta J, Oliveres J, Piñol J, Lloret F (2010) Feedbacks between fuel reduc-
tion and landscape homogenisation determine fire regimes in three Mediterranean areas. For
Ecol Manag 259(12):2366–2374
Madrigal J, Hernando C, Martínez E, Guijarro M, Díez C (2005) Regeneración post-incendio de
Pinus pinaster Ait. en la Sierra de Guadarrama (Sistema Central, España): modelos descrip-
tivos de los factores influyentes en la densidad inicial y la supervivencia. Inv Agr Sist Recur
For 14:36–51
Martín-Alcón S, Coll L (2016) Unraveling the relative importance of factors driving post-fire
regeneration trajectories in non-serotinous Pinus nigra forests. For Ecol Manag 361:13–22.
https://doi.org/10.1016/j.foreco.2015.11.006
Martín-Alcón S, Coll L, Ameztegui A (2015) Diversifying sub-Mediterranean pinewoods with oak
species in a context of assisted migration: responses to local climate and light environment.
App Veg Sci 19:254–266
Matthews JD (1991) Silvicultural Systems. Oxford University Press, Oxford
Mayor AG, Valdecantos A, Vallejo VR, Keizer JJ, Bloem J, Baeza J, Gonzalez-Pelayo O, Machado
AI, de Ruiter PC (2016) Fire-induced pine woodland to shrubland transitions in southern
Europe may promote shifts in soil fertility. Sci Total Environ 573:1232–1241. https://doi.
org/10.1016/j.scitotenv.2016.03.243
McDowell NG (2011) Mechanisms linking drought, hydraulics, carbon metabolism, and vegeta-
tion mortality. Plant Physiol 155(3):1051–1059. https://doi.org/10.1104/pp.110.170704
McDowell N, Pockman WT, Allen CD, Breshears DD, Cobb N, Kolb T, Plaut J, Sperry J, West A,
Williams DG, Yepez EA (2008) Mechanisms of plant survival and mortality during drought:
why do some plants survive while others succumb to drought? New Phytol 178(4):719–739
Midgley JJ, Lawes MJ, Chamaillé-Jammes S (2010) Savanna woody plant dynamics: the role of
fire and herbivory, separately and synergistically. Aust J Bot 58:1–11
Morales-Molino C, Tinner W, García-Antón M, Colombaroli D (2017) The historical demise of
Pinus nigra forests in the Northern Iberian Plateau (South-Western Europe). J Ecol 105:634–646
Moreno JM, Oechel WC (1991) Fire intensity and herbivory effects on postfire resprouting of
Adenostoma fasciculatum in southern California chaparral. Oecologia 85(3):429–433. https://
doi.org/10.1007/bf00320621
Moreno JM, Zuazua E, Pérez B, Luna B, Velasco A, Resco De Dios V (2011) Rainfall patterns
after fire differentially affect the recruitment of three Mediterranean shrubs. Biogeosciences
8(12):3721–3732
Noble IR, Slatyer RO (1980) The use of vital attributes to predict successional changes in plant
communities subject to recurrent disturbances. Vegetatio 43:5–21
Nunes MCS, Vasconcelos MJ, Pereira JMC, Dasgupta N, Alldredge RJ, Rego FC (2005) Land
cover type and fire in Portugal: do fires burn land cover selectively? Landsc Ecol 20(6):661–
673. https://doi.org/10.1007/s10980-005-0070-8
Odion DC, Moritz MA, Dellasala DA (2010) Alternative community states main-
tained by fire in the Klamath Mountains, USA. J Ecol 98(1):96–105. https://doi.
org/10.1111/j.1365-2745.2009.01597.x
Ordóñez JL, Retana J (2004) Early reduction of post-fire recruitment of Pinus nigra by post-­
dispersal seed predation in different time-since-fire habitats. Ecography 27:449–458
Ordóñez JL, Retana J, Espelta JM (2005) Effects of tree size, crown damage, and tree location
on post-fire survival and cone production of Pinus nigra trees. For Ecol Manag 206:109–117.
https://doi.org/10.1016/j.foreco.2004.10.067
Ordoñez J, Molowny-Horas R, Retana J (2006) A model of the recruitment of Pinus nigra from
unburned edges after large wildfires. Ecol Model 197:405–417
Parra A, Moreno JM (2017) Post-fire environments are favourable for plant functioning of seeder
and resprouter Mediterranean shrubs, even under drought. New Phytol. https://doi.org/10.1111/
nph.14454
Pausas JG (2015) Evolutionary fire ecology: lessons learned from pines. Trends Ecol Evol 20:318–
324. https://doi.org/10.1016/j.tplants.2015.03.001
152 8 Forest Succession, Alternative States, and Fire-Vegetation Feedbacks

Pausas JG, Moreira B (2012) Flammability as a biological concept. New Phytol 194(3):610–613.
https://doi.org/10.1111/j.1469-8137.2012.04132.x
Pausas JG, Ouadah N, Ferran A, Gimeno T, Vallejo R (2003) Fire severity and seedling establish-
ment in Pinus halepensis woodlands, eastern Iberian Peninsula. Plant Ecol 169:205–213
Pausas JG, Blade C, Valdecantos A, Seva JP, Fuentes D, Alloza JA, Vilagrosa A, Bautista S, Cortina
J, Vallejo R (2004) Pines and oaks in the restoration of Mediterranean landscapes of Spain: new
perspectives for an old practice – a review. Plant Ecol 171:209–220
Pausas J, Llovet J, Rodrigo A, Vallejo R (2008) Are wildfires a disaster in the Mediterranean
basin? – A review. Int J Wildland Fire 17:713–723
Peguero-Pina JJ, Sancho-Knapik D, Martín P, Saz MÁ, Gea-Izquierdo G, Cañellas I, Gil-­
Pelegrín E (2015) Evidence of vulnerability segmentation in a deciduous Mediterranean oak
(Quercus subpyrenaica E. H. del Villar). Trees 29(6):1917–1927. https://doi.org/10.1007/
s00468-015-1273-5
Peñuelas J, Sardans J, Filella I, Estiarte M, Llusià J, Ogaya R, Carnicer J, Bartrons M, Rivas-­
Ubach A, Grau O, Peguero G, Margalef O, Pla-Rabés S, Stefanescu C, Asensio D, Preece
C, Liu L, Verger A, Rico L, Barbeta A, Achotegui-Castells A, Gargallo-Garriga A, Sperlich
D, Farré-Armengol G, Fernández-Martínez M, Liu D, Zhang C, Urbina I, Camino M, Vives
M, Nadal-Sala D, Sabaté S, Gracia C, Terradas J (2017) Assessment of the impacts of cli-
mate change on Mediterranean terrestrial ecosystems based on data from field experiments and
long-term monitored field gradients in Catalonia. Environ Exp Bot. https://doi.org/10.1016/j.
envexpbot.2017.05.012
Pratt RB, Jacobsen AL, Ramirez AR, Helms AM, Traugh CA, Tobin MF, Heffner MS, Davis
SD (2014) Mortality of resprouting chaparral shrubs after a fire and during a record drought:
physiological mechanisms and demographic consequences. Glob Chang Biol 20(3):893–907.
https://doi.org/10.1111/gcb.12477
Quero JL, Gómez-Aparicio L, Zamora R, Maestre FT (2008) Shifts in the regeneration niche of
an endangered tree (Acer opalus ssp. granatense) during ontogeny: using an ecological concept
for application. Basic Appl Ecol 9(6):635–644. https://doi.org/10.1016/j.baae.2007.06.012
Quintana JR, Cruz A, Fernández-González F, Moreno JM (2004) Time of germination and estab-
lishment success after fire of three obligate seeders in a Mediterranean shrubland of Central
Spain. J Biogeogr 31(2):241–249
Resco de Dios V, Fischer C, Colinas C (2007) Climate change effects on Mediterranean forests and
preventive measures. New For 33:29–40
Resco de Dios V, Weltzin JF, Sun W, Huxman TE, Williams DG (2014) Transitions from grassland
to savanna under drought through passive facilitation by grasses. J Veg Sci 25:937–946
Resco V, Ewers BE, Sun W, Huxman TE, Weltzin JF, Williams DG (2009) Drought-induced
hydraulic limitations constrain leaf gas exchange recovery after precipitation pulses in the C3
woody legume, Prosopis velutina. New Phytol 181:672–682
Retana J, Espelta JM, Habrouk A, Ordóñez JL, de Solà-Morales F (2002) Regeneration patterns
of three Mediterranean pines and forest changes after a large wildfire in northeastern Spain.
Écoscience 9:89–97
Richard F, Selosse MA, Gardes M (2009) Facilitated establishment of Quercus ilex in shrub-­
dominated communities within a Mediterranean ecosystem: do mycorrhizal partners matter?
FEMS Microbiol Ecol 68(1):14–24. https://doi.org/10.1111/j.1574-6941.2009.00646.x
Rivas-Martínez and co-authors (2005) Mapa de series, geoseries y geopermaseries de vegetación
de España. Phytosociological Research Center, Madrid
Rodrigo A, Quintana V, Retana J (2007) Fire reduces Pinus pinea distribution in the North-Eastern
Iberian Peninsula. Ecoscience 14:23–30
Ruiz-Labourdette D, Nogués-Bravo D, Ollero HS, Schmitz MF, Pineda FD (2012) Forest composi-
tion in Mediterranean mountains is projected to shift along the entire elevational gradient under
climate change. J Biogeogr 39(1):162–176. https://doi.org/10.1111/j.1365-2699.2011.02592.x
References 153

Sagra J, Moya D, Plaza-Álvarez P, Lucas-Borja M, Alfaro-Sánchez R, De Las Heras J, Ferrandis


P (2017) Predation on early recruitment in Mediterranean forests after prescribed fires. Forests
8(7):243. https://doi.org/10.3390/f8070243
Santana VM, Baeza MJ, Marrs RH, Vallejo VR (2010) Old-field secondary succession in SE Spain
: can fire divert it ? Plant Ecol 211:337–349. https://doi.org/10.1007/s11258-010-9793-y
Santana VM, Baeza MJ, Maestre FT (2012) Seedling establishment along post-fire succession in
Mediterranean shrublands dominated by obligate seeders. Acta Oecol 39:51–60. https://doi.
org/10.1016/j.actao.2011.12.001
Simard S, Martin K, Vyse A, Larson B (2013) Meta-networks of fungi, fauna and flora as agents
of complex adaptive systems. In: Messier C, Puettmann KJ, Coates KD (eds) Managing forests
as complex adaptive systems: building resilience to the challenge of global change. Routledge,
New York, pp 133–164
Skelton RP, Brodribb TJ, McAdam SAM, Mitchell PJ (2017) Gas exchange recovery following
natural drought is rapid unless limited by loss of leaf hydraulic conductance: evidence from an
evergreen woodland. New Phytol 215(4):1399–1412. https://doi.org/10.1111/nph.14652
Smith SE, Read DJ (2008) Mycorrhizal symbiosis, 3rd edn. Academic Press, New York
Tapias R, Climent J, Pardos JA, Gil L (2004) Life histories of Mediterranean pines. Plant Ecol
171(1–2):53–68
Taudière A, Richard F, Carcaillet C (2017) Review on fire effects on ectomycorrhizal symbiosis, an
unachieved work for a scalding topic. For Ecol Manag 391:446–457. https://doi.org/10.1016/j.
foreco.2017.02.043
Thanos CA, Daskalakou EN (2000) Reproduction in Pinus halepensis and P. brutia. In: Néeman
G, Trabaud L (eds) Ecology, biogeography and management of Pinus halepensis and P. brutia
forest ecosystems in the Mediterranean Basin. Backhuys Publishing, Leiden, pp 79–90
Tomao A, Bonet JA, Martínez de Aragón J, de-Miguel S (2017) Is silviculture able to enhance
wild forest mushroom resources? Current knowledge and future perspectives. For Ecol Manag
402:102–114. https://doi.org/10.1016/j.foreco.2017.07.039
Torrecillas E, Alguacil MM, Roldán A (2012) Host preferences of arbuscular mycorrhizal fungi
colonizing annual herbaceous plant species in semiarid Mediterranean prairies. Appl Environ
Microbiol 78(17):6180–6186. https://doi.org/10.1128/AEM.01287-12
Torres I, Pérez B, Quesada J, Viedma O, Moreno JM (2016) Forest shifts induced by fire and
management legacies in a Pinus pinaster woodland. For Ecol Manag 361:309–317. https://doi.
org/10.1016/j.foreco.2015.11.027
Trabaud L (1984) Fire adaptation strategies of plants in the french Mediterranean area. In: Margaris
NS, Arianoustou-Faraggitaki M, Oechel WC (eds) Being alive on land, Tasks for vegetation
science, vol 13. Springer, Dordrecht, pp 63–69
Trabaud L, Grosman J, Walter T (1985) Recovery of burnt Pinus halepensis mill. Forests.
I. Understorey and litter phytomass development after wildfire. For Ecol Manag 12:269–277
Vallejo VR, Alloza JA (1998) The restoration of burned lands: the case of eastern Spain. In:
Moreno JM (ed) Large forest fires. Backhuys Publishing, Leiden, pp 91–108
Vieira DCS, Fernandez C, Vega JA, Keizer JJ (2015) Does soil burn severity affect the post-fire
runoff and interrill erosion response? A review based on meta-analysis of field rainfall simula-
tion data. J Hydrol 523:452–464
Vilà-Cabrera A, Martínez-Vilalta J, Vayreda J, Retana J (2011) Structural and climatic deter-
minants of demographic rates of scots pine forests across the Iberian Peninsula. Ecol Appl
21:1162–1172. https://doi.org/10.1890/10-0647.1
Wilkinson DM, Sherratt TN (2016) Why is the world green? The interactions of top–down and
bottom–up processes in terrestrial vegetation ecology. Plant Ecol Diver 9(2):127–140. https://
doi.org/10.1080/17550874.2016.1178353
Woods SR, Archer SR, Schwinning S (2014) Seedling responses to water pulses in shrubs with
contrasting histories of grassland encroachment. PLoS One 9:e87278. https://doi.org/10.1371/
journal.pone.0087278.g001
Chapter 9
Pyrophysiology and Wildfire Management

Abstract In this chapter we further discuss how can pyrophysiology assist us with
wildfire management. The overall objective of wildfire management is to reduce the
risk to human lives and properties while maintaining the role of fire as a natural
process in ecosystems. Wildfire management thus seeks to diminish fuel loads
beyond the critical levels that could lead to extremely large fires and endanger civil
security. Pyrophysiology may help to achieve this goal by developing economic and
sustainable land management alternatives that lead to fire-resistant landscapes.
Pyrophysiology also informs on which are the most appropriate fuel reduction treat-
ments (e.g., mechanical vs prescribed burn) or in the ways to achieve fuel disconti-
nuities (e.g., green firebreaks) based on vegetation physiological responses. It is
important to note that pyrophysiology may only provide a limited information on
extremely large fire behavior because fuel traits affecting flammability often act
over small scales (e.g., leaf). Fire behavior models similarly show limited predictive
capabilities over large fires, and consideration of physiological mechanisms may
render more realistic predictions of diurnal, seasonal, and annual fuel dynamics but
to limited increases in model fit. We end this chapter by discussing how pyrophysi-
ology is of relevance for deciding the most appropriate post-fire restoration actions.

9.1 Objectives of Wildfire Management

Wildfires are a natural component of many ecosystems worldwide, and they have
been present on Earth for over 400 mya (Chap. 3). Before defining any management
action related to wildfires, we must decide the goals of such management plan. Fire
suppression has often been the sole objective in wildfire management, and we have
so far sought to put out a fire as soon as possible after it started. This paradigm has
been changing in the last few years, at least in some parts of the world, after observ-
ing that wildfires are essential ecosystem components in many parts of the world
and after appreciating the fact that fire suppression enhances large fire activity (the
so-called firefighting trap (Fernandes et al. 2020)).
Managing wildfires should seek to achieve fire-smart landscapes, that is, to
ensure citizens’ safety as well as of maintaining ecosystem health and, where appli-

© Springer Nature Switzerland AG 2020 155


V. Resco de Dios, Plant-Fire Interactions, Managing Forest Ecosystems 36,
https://doi.org/10.1007/978-3-030-41192-3_9
156 9 Pyrophysiology and Wildfire Management

cable, allowing economic development sensu lato (i.e., including non-woody forest
products and other goods and services). The first objective of wildfire management
thus lies in reducing the risk for life and property and particularly valuable ecosys-
tems. Wildfires are becoming an increasing civil protection problem: we can count
383 fatalities as a direct result of wildfires between 2000 and 2016 in Spain,
Portugal, Greece, and Sardinia (Molina-Terrén et al. 2019), a number comparable to
the number of terrorism victims during that period.
We must ensure people’s safety as well as ecosystem health. Plant species show a
number of physiological adaptations to fire (Chap. 4), and, in particular, their regenera-
tion may be linked to a certain wildfire regime. Wildfires are thus an essential ecological
process, and we should strive to maintain such pyrodiversity through land management.
Wildfire management is a double-edged sword. On the one hand, we must ensure
the safety of social, economic, and ecological systems. On the other hand, we some-
times need to introduce fire back as a management tool.
A result of fire suppression has been a fire “deficit.” Burnt area is currently lower
than what it should be considering current climate and oxygen concentrations. This
creates a series of problems, such as a surplus of biomass that is then available to
burn in large fires. Even if burnt area has declined steadily in Europe for the past few
decades, the intensity with which fires burn has increased. Such increase in fire
intensity likely results in biomass surplus. But how did we get here? We will now
examine the evolution of the fire problem before trying to understand its management.
We begin by explaining the fire generations concept. This concept was developed
for Southern Europe, although the problems associated with fourth-, fifth-, and
sixth-generation fires are common across many areas with wildfire problems glob-
ally. Understanding the different fire generations is useful to understand how the fire
problem has developed. Afterwards, we will introduce the principles underlying
landscape management for reducing the likelihood of large wildfires through the fire
types concept. After this introduction to management, we will describe how pyro-
physiology may inform for selecting the most appropriate fuel treatment: prescribed
burns, mechanical treatments, or green firebreaks. We will then describe the limita-
tions of pyrophysiology in predicting landscape levels of flammability, the problems
currently facing fire behavior models, and how can we achieve more realistic repre-
sentations of fuel dynamics in fire models by addition of ecophysiological informa-
tion. In the last section, we will call for more studies jointly using physiological
research to better understand how to manage and restore post-fire areas.

9.2 Fire Generations

The fire generations concept (Table 9.1) originated to describe the changes in fire
behavior since the rural exodus in Mediterranean Southern Europe (Castellnou and
Miralles 2010; Costa et al. 2011). It presents a general description of how fire inten-
sity, fire prevention, and firefighting strategies change as more time is allowed for
fuel build-up. First-generation fires occur after a brief period of fuel accumulation
(2–15 years). They are thus driven by the increase in wildland fuel continuity, and
9.2 Fire Generations 157

Table 9.1 Different fire generations. Information for generations 1–5 from Costa et al. (2011)
Generation Defined by Fire behavior Fire prevention
First Horizontal fuel Surface fires of medium Linear infrastructures, water
continuity intensity points, and improved access
Second Increased fuel Surface fires with high Shortening of response times.
build-up leading to intensity advancing with Linear infrastructures may be
high rate of spread spot fires overcome by spot fires
Third Vertical fuel High-intensity crown fires Develop confinement
continuity strategies: areas with lower
fuel loads where fuel
intensity declines
Fourth Crossing wildland-­ Independent crown fires Pyrogardening, fuel reduction
urban interface burn across the continuous treatments in urban areas to
(WUI) fuel between wildlands and create discontinuities
urban areas
Fifth Simultaneous fronts Simultaneous crown fires Considering fire into land
crossing the WUI concomitantly affecting management
different areas. Very rapid
and virulent
Sixth Pyroconvection Extreme fire behavior over Drastic changes in landscape
events tens or hundreds of structure that decrease
thousands of hectares. dramatically fuel loads and
Extremely rapid and erratic inhibit pyroconvection
fire behavior activity

surface fires may burn at low to mid intensity. These fires are often controlled by
linear infrastructures such as firebreaks, increased presence of water points, and
granting better access through roads.
Second-generation fires occur after 10–30 years of fuel accumulation1, and they
start to present control problems. Here, the fuel build-up is high enough to carry
fires with large intensities and high rates of spread. Consequently, linear firebreaks
may not be enough to stop a fire as the flames may be able to “jump” over such
infrastructures. Firefighting lies in short response times, such that suppression
actions start before the fire has reached large intensities. In turn, this requires an
extensive system for the detection of new fires and a large distribution of fire
stations.
Third-generation fires are characterized by high-intensity crown fire behavior.
They may appear after 30–50 years of land abandonment. These fires generate con-
vective columns and large spot fires, such that there is little opportunity for direct
attacks. Firefighting strategies are based upon confinement and increased use of fire
as a suppression tool, along with other measures that allow confinement.
Fourth-generation fires may start and end at the wildland-urban interface (WUI),
and fifth-generation fires refer to the occurrence of simultaneous large fires at the
WUI, which present a very rapid and virulent behavior. These fires result from the

1
Note that fuel accumulation times refer to the productivities from Mediterranean S Europe. They
may thus vary for other regions if their productivities are different.
158 9 Pyrophysiology and Wildfire Management

continuity in fuel between the urban area and the surrounding wildland. Such fires
may burn as crown fires and generate major security problems. Prevention actions
include pyrogardening, the breakage of fuel continuity between the wildland and
the urban area, constructing fireproof buildings, and overall protecting houses
from fire.
Sixth-generation fires are the most destructive fire types we have experienced to
date. They are able to consume 10,000 ha or more in just 1 h (Fig. 9.1) (Castellnou
et al. 2018; Ribeiro et al. 2020). Fuel accumulated over more than 50 years has a
very large potential energy. Under fire, the energy released is so high that it may
give rise to pyroconvective activity. The convective column may then lead to the
formation of pyrocumulus (pyroCu) or pyrocumulonimbus (pyroCb). These may
collapse in the evening, as temperatures lower and atmospheric humidity declines,
or when the front reaches areas with less available fuel, and the result is extremely
erratic and rapid fire behavior.
8000 10000 12000 14000
)1
Fire growth rate (ha h
6000
4000
2000
0

9 12 15 18 21 24 3 5 7 9 12 15
Time (h)

Fig. 9.1 Example of a sixth-generation fire starting on October 15, 2017, in Central Portugal.
These are unprecedented rates of fire spread. The objective of wildfire management is to seek
landscapes where these fires cannot occur because of fuel limitations. Studies that seek to under-
stand this type of fire behavior from physiological traits related to flammability or from fire behav-
ior modeling have difficulties in reproducing these results. (Data from Castellnou et al. (2018))
9.3 Fire Spread Patterns and Preventive Measures 159

9.3 Fire Spread Patterns and Preventive Measures

In order to understand how ecophysiology may help in wildfire management, it is


necessary to describe at least the basic principles underlying wildfire prevention.
Fires may be categorized based upon their pattern of spread as topographic or wind-
or plume-driven fires (Duane et al. 2015). Topographic fires spread following hydro-
logic basins in combination with local, orographic winds and the differential heating
and cooling that occurs within slopes during day and night. Wind-driven fires follow
the prevailing wind direction, and their spreading interacts with ridge orientation in
mountainous areas. Wind-driven fires may burn at high intensity because of the
increased aeration (forced convection), and they may propagate quickly by spotting.
Plume-driven fires are the most dangerous and unpredictable fire types. They are
driven by the convective currents created by the fire, which are lofted and thus gen-
erate wind drafts on the surface that may lead into massive spotting and very rapid
and erratic fire spread. These three different fire types should not be viewed as
antagonistic, and different combinations of fire types are possible.
Pau Costa, a forest engineer and firefighter, pioneered the notion of managing
wildfires depending on the fire type (Costa et al. 2011). That is, these different wild-
fire types propagate very differently across the landscape, and they are also driven
by very different factors. Consequently, wildfire management and prevention should
be tailored to the specific set of fire traits that may occur within a region. In fact,
topographic and wind-driven fires are jointly driven by unmanageable factors, such
as topographic or climatic variables, and by landscape scale factors, such as the
degree of landscape heterogeneity and the different land covers present. However,
convective fires are driven mostly by wildland fuel factors related to the total bio-
mass and structure (Duane et al. 2015). Given these different drivers of the pattern
of spread, we may thus devise different managing options for different fire types.
The original fire prevention strategy consisted of developing firebreaks, aiming
at defending a stand from incoming fires by creating an anchor point to be used by
fire brigades. The increase in fire intensity in the last few decades creates problems
with this strategy as ambers and firebrands are capable of jumping over such linear
infrastructures, hence the need to develop management plans over larger areas. For
instance, sixth-generation plume-driven fires are effectively unstoppable once they
develop. The only prevention strategy is to manage the landscape at regional scales
and to diminish available fuel loads beyond critical intensities that may give rise to
the formation of pyroCu and pyroCb.
A possible strategy to achieve this goal, at regional scales, is to first define areas
which are homogeneous from the viewpoint of the fire regime (Costa et al. 2011).
These are areas where vegetation and past fire regime are similar and we can char-
acterize the fire type that we can expect to occur in that area. This is a useful clas-
sification because fire spread patterns are not random over a region. On the contrary,
the fire regime within an area is defined by the interactions between fuel, topogra-
phy, and meteorology.
160 9 Pyrophysiology and Wildfire Management

After dividing a region into homogeneous fire regime areas, we identify strategic
management areas (SMAs) over landscapes. Rather than selecting randomly the
areas where we will conduct fire prevention measures, it is preferable to act over
SMAs, critical points where a fire could spread into new areas and increase its
potential if not controlled.
In wildfire prevention we are overall seeking to limit both fire intensity and
spread, to confine ignition and to improve access to the fire. For areas dominated by
topographic fires, this includes managing, for instance, the bottom of the ravines, as
this is an area where the fire could enter into a different basin. For areas dominated
by wind-driven fires, in contrast, we will be seeking to manage areas at the top of
the ravine, as this is an area where fire potential could multiply if the fire escapes
into new ravines. The reader is referred to the original publication for further details
on this (Costa et al. 2011) and taking in mind that this approach has been developed
for an area in Southern Europe. We acknowledge there is a wide view on how to
develop wildfire prevention actions, and a full review would be out of scope. Here
it will suffice to keep in mind that different wildfire types require different fire pre-
vention treatments.
It is worth noting that the efficiency of such management approach still needs to
be quantified. The efficiency of fire prevention actions declines as the annual area
burnt is smaller. As a general rule of thumb, it is considered that it is necessary to
treat three unit areas in order to diminish burnt area by one unit area (Boer et al.
2009). The relative proportion of land area that needs to be treated increases as the
percent annual burnt area (relative to total wildland area) declines (Price et al. 2015).
This is because wildfires are a rare occurrence. It is unlikely that a fire will cross an
area that has been previously treated, unless such fire prevention systems are exten-
sive over the land (Boer et al. 2015). The smaller the percent burnt area, the less
likely that it will be intercepted by a fire, and, in consequence, a larger area needs to
be treated. Managing SMAs is thus important to increase the efficiency of fire pre-
vention treatments.
If a fire intercepts one of the SMAs that has been properly treated, that does not
imply that the wildfire will automatically be extinguished. It simply serves as an
opportunity to decline the fire intensity and to increase the chances of firefighters
controlling the fire. As fuel build-up increases, the area that needs to be treated
increases because fire intensity keeps increasing, with some estimates indicating
that 14% of the land must be treated in order to diminish the chances of a large fire
in a Mediterranean county (Alcasena et al. 2017).
After establishing a SMA, maintenance is essential. After a fuel reduction treat-
ment has been implemented, it is only a matter of time that fuel loads return to simi-
lar levels as those occurring originally. Time to recover differs as a function of
treatment intensity, site productivity, and species growth rates, but they often lie
between 5 and 15 years (Soler Martin et al. 2017; Agee and Skinner 2005).
The fire problem, at least in Europe, has developed after a change in our socio-
economic model where the traditional energy source, firewood in the wildland, has
been replaced by fossil fuels. We thus cannot expect a quick fix to such problem,
and, unless we transform our landscapes, the frequency of sixth-generation fires
9.4 Pyrophysiology and Wildfire Prevention 161

might increase. Fortunately, advances in applied ecophysiology allow for the gen-
eration of economic alternatives that may help in the creation of less flammable
landscapes, as we discuss in the next section.

9.4 Pyrophysiology and Wildfire Prevention

There are different types of wildfire prevention strategies that may be implemented
within land planning. The ideal measure would be the creation of landscapes that
are naturally not prone to large-scale fires. That is, wildfires are a native component
of many ecosystems worldwide, and we should not seek to completely eradicate
them from our landscapes. We should learn to coexist with fire, and this means that
we need to establish a fire prevention system that maintains citizens’ safety but also
one where wildfires of adequate scale and intensity occur (Moritz et al. 2014). The
solution to this problem thus lies in modifying landscape features such that large
fires are very difficult to occur due to the degree of landscape heterogeneity and
because of the reduced fuel loads. In other words, we must seek to diminish the
potential for large and extreme fires but maintain small fires as natural ecosystem
processes.
The ideal wildfire prevention strategy would thus be a shift in the economic para-
digm that allows for sustainable land management to be profitable and, conse-
quently, where fuel loads are maintained below the threshold required for the
development of large fires that may be beyond extinction capacity. However this is
not always possible, and fuel reduction treatments in SMAs along with other strate-
gies that hinder fire propagation such as the so-called green firebreaks will be
necessary.
Here we begin by discussing recent advances in ecophysiology that increase
wildland profitability in low productivity lands such as the Mediterranean and that
may help in transforming our landscapes. Next we discuss the mechanisms explain-
ing how silvicultural treatments diminish water stress in wildland areas and, conse-
quently, diminish fire risk. Then we will discuss the ecophysiology behind fuel
reduction treatments in SMAs. One of the most contentious fuel reduction treat-
ments has been prescribed burning. There are many reasons explaining why some
land managers are reluctant to use fire in landscape, and one of them is the fear that
prescribed burning may damage the remaining trees. We will thus examine tree
ecophysiological responses to prescribed burning. We will also provide examples of
how to use ecophysiology to decide the fuel reduction treatment (e.g., mechanical
vs prescribed fire) depending upon the physiological traits of the species present.
Finally, we will discuss the development of fuel greenbreaks, which has been pro-
posed as a strategy to hinder fire spread in some areas.
162 9 Pyrophysiology and Wildfire Management

9.4.1  evelopment of Economically Viable Landscapes


D
Through Ecophysiology

Advancements in ecophysiology have increased the range of economic opportuni-


ties that make fire prevention treatments feasible. Pinus halepensis is the most com-
mon pine species in the Mediterranean basin. This species is not very attractive
economically because it shows limited yield, a problem that is further exacerbated
because it often grows over poor soils, and most forests are thus abandoned.
Pinus pinea, a tree that produces edible pine seeds, grows mainly on acidic soils,
but it can be grafted on P. halepensis roots (Piqué et al. 2017). Such graftings may
be a possibility to promote forest management in the area such that fuel loads are
reduced. At the moment of writing, pine seeds are priced at 40 € kg−1, although this
productivity is currently threatened by pests infestations.
Grafting is possible with pines but also with other forest species, such as walnuts
(Juglans regia), and the opportunities for using this technique to foster the develop-
ment of economically profitable landscapes thus deserves further attention.
The productivity of such plantations, at least at the seedling stage, may be further
exacerbated by the development of soil conditioners and mulching products that
hold up to 400 times their volume in water. These mulches could increase seedling
survival and growth even in semiarid environments (Coello et al. 2018).
There is also a renewed interest in using biomass as an energy sources. According
to the last IPCC Assessment Report (IPCC 2014), technologies such as bioenergy
with carbon capture and storage, which produces energy from biomass burning and
stores its emissions, will be necessary to achieve negative emissions in the near
future. Wood chips coming from woody species may be another alternative to fund
the existence of fire-resilient landscapes.
Ecophysiology informs of the growth requirements and niche tolerances of dif-
ferent species. This will be particularly relevant under the novel environmental con-
ditions that will arise from global warming, as we can expect shifts in species
distributions and restoration efforts that include new tree plantations. Understanding
ecophysiological traits will thus be useful for predicting the species most suitable
under novel environmental conditions as well as for biomass uses.

9.4.2 Silvicultural Treatments Diminishing Water Stress

We have previously described how large wildfires in the Mediterranean are the
result of land abandonment, and it is often rare to observe large wildfires building
up in properly managed landscapes (Fig. 9.2). Fires only seldom occur in areas with
silvicultural interventions that break the fuel ladder and also crown overlap. Under
such stands, lower tree densities lead to a lower degree of water stress and, conse-
quently, to higher fuel moistures and to lower dead to live fuel ratios. We also
observe much lower surface fuel loads, which difficult initial fire spread. Lack of
9.4 Pyrophysiology and Wildfire Prevention 163

Fig. 9.2 Pine plantation


with reduced fuel loads
due to thinning and
prescribed burning in Plans
de Sant Joan, Tarragona,
Spain

overlap in the crowns diminishes the canopy bulk density (CBD), further hindering
crown fire propagation.
The underlying mechanism for higher fuel moistures and decreased dead to live
ratios is that thinning and other silvicultural treatments that reduce competition
increase the degree of water available to the remaining individuals (see Chap. 7).
Indeed it is often observed a higher water use at the leaf level, indicated by higher
stomatal conductance, after thinning (Sala et al. 2005; Moreno-Gutierrez et al.
2011). It is interesting to note that, despite an amelioration in tree water status, the
effects on carbon uptake seem to depend on water availability. Increases in post-­
thinning net photosynthesis and growth have been documented for mesic (Sala et al.
2005; Zausen et al. 2005), but not always for xeric landscapes (Moreno-Gutierrez
et al. 2011).
However, such treatments must be performed carefully. If inter-crown spacing is
too small, then an intervention may be necessary in the near future, but, if they are
too large, surface radiation loads may be high enough to allow for the growth of
understory vegetation, and, in some cases, litter humidity may also decline and
wind speeds might increase.
Understanding how often these treatments should be applied is a function of both
site productivity and physiological traits such as growth rates or canopy a­ rchitecture.
For pine stands, it is often estimated that it takes about 15–20 years to recover CBD
164 9 Pyrophysiology and Wildfire Management

to the values before the intervention, but other studies have determined that a full
CBD recovery may be observed only 7 years after the application (Jiménez et al.
2016; Soler Martin et al. 2017). It is also important to determine the best time of the
season to perform such treatments, as dead biomass may increase after the treatment
and the fire risk may be higher in the dry season.

9.4.3 Fuel Reduction Treatments

One of the most common approaches for fire prevention lies in diminishing surface
fuel loads in SMAs, as described above. There are different techniques to develop
this, and prescribed burning is one of them (Fig. 9.2). That is, we may decide to
conduct a controlled fire in order to get rid of the understory vegetation, to break the
fuel ladder or to diminish litter and surface fuel loads, as an example. Alternatively,
surface fuels may be mechanically treated.

9.4.3.1 Impacts of Prescribed Burns on Tree Physiology

Prescribed burns may impact the physiology of surviving trees, but the effects
depend on interactions between burn intensity, competition release, and insect
attacks. Studies conducted in pine trees often report increases in stomatal conduc-
tance and water use after the burn, indicating higher water availability due to com-
petition removal (Valor et al. 2018). Short-term declines in radial growth may occur
during the first year after the burn (Sparks et al. 2017), particularly in individuals
with a significant portion of their crown scorch (Valor et al. 2018). However, long-­
term growth effects are often positive because of competition removal and because
of a lack of negative effects over tree hydraulics and anatomy under the low fire
intensities typical of prescribed burns (Battipaglia et al. 2016).
Additionally, low-intensity fires may induce changes in resin yield such that
conifers may be less susceptible to consequent insect attacks. This has been observed
in ponderosa pine, for instance, where resin duct area increased significantly for the
first 3–5 years after the burn (Sparks et al. 2017). Resin duct area resumed its pre-­
bun values after this period (Hood et al. 2015). This response is not universal across
pine species since, for instance, no changes in resin duct area or yield were observed
in P. pinaster (Rodríguez-García et al. 2018). Other studies have found that any
potential increase in tree defenses may not be enough to overcome increased insect
attacks when the burn is very intense (Perrakis et al. 2011). Apart from changes in
resin duct area, shifts in resin chemistry that affect communication across mountain
pine beetle have also been reported. Such changes in resin chemistry are thought to
indirectly reduce beetle attack success by interfering with insect communication
(Hood et al. 2016). Overall, the effects of prescribed burn over tree defenses are
unclear, and current studies are limited to conifers. At any rate, current evidence
9.4 Pyrophysiology and Wildfire Prevention 165

indicates that prescribed burning either increases or does not change tree suscepti-
bility to pest attacks, while reports of increased susceptibility are rare.

9.4.3.2 Prescribed Burn vs Mechanical Treatment

Prescribed burning is economically beneficial for the treatment of large areas, as it


is inexpensive over such vast areas. It offers an opportunity to train fire crews in the
use of fire as a management tool, which, for instance, may be necessary to develop
backfires during suppression efforts. Furthermore, it allows for the reintroduction of
fire into ecosystems, meaning that it can serve to re-establish the natural fire regime
within an area (with caveats, as prescribed burns are often conducted out of the main
fire season). However, prescribed burns have the setback of emitting higher green-
house gas emissions relative to mechanical treatments, and they also emit larger
amounts of particular matter and other pollutants. Smoke from prescribed burning
may also alarm citizens living near the managed stands.
Regardless of the different advantages and problems associated with prescribed
fires, from an ecophysiological perspective, they are particularly advised for the
treatment of resprouting vegetation. Burning diminishes biomass loads and subjects
the plants to a thermal stress, which hinders resprouting to a greater degree than
mechanical treatments alone. That is, mechanical treatments do not induce a ther-
mal stress over the plant buds, and resprouting may be more vigorous than under
prescribed fire. If such prescribed fire is applied after a period of strong drought,
resprouting will be further hindered.
On the contrary, prescribed burnings are discouraged in areas dominated by
seeder species with seed banks. Germination after fire may increase (directly or
indirectly; see Chaps. 4, 7, and 8), relative to the application of a mechanical treat-
ment. In this case, the understory vegetation may recover more quickly after fire
than after a mechanical treatment such as mastication.
Mastication alters community structure, relative to both controls with and with-
out fire. Annual and ruderal species often benefit from mastication, as these are
species with very fast establishment rates after disturbance. This sometimes leads to
an overall higher diversity of annual plants after mastication. In a site from
California, it was observed that non-native species were particularly favored by the
mastication treatment (Brennan and Keeley 2017). However, the cover and density
of woody shrubs often decline, relative to burnt plots, due to the negative effects of
fire on resprouting (relative to biomass elimination alone). Other studies, however,
have observed that mastication does not favor the establishment of invasives
(Fernández et al. 2019), and more research is thus needed to better understand the
impacts of this technique.
Mechanical treatments and prescribed burning are sometimes accompanied by
thinning. Studies comparing the effects thinning alone or with burning indicate that
the positive effect of prescribed burning on growth, which is mediated by increased
water availability, often exerts little additional benefit to that already provided by
thinning. Thus, while potentially advantageous for fuel reduction, additional stress
166 9 Pyrophysiology and Wildfire Management

reductions from concurrent application of thinning and burn are often scarce (Hood
et al. 2016; Sala et al. 2005).

9.5 Green Firebreaks

Another infrastructure that has sometimes been used to control fires is green fire-
breaks (Cui et al. 2019). This is essentially a wall of tightly packed trees that may
be effective in preventing the spread of low-intensity fires. The tree species used are
those with traits of low leaf-level flammability, such as high moisture content and
ash content, along with other properties such as low dead to live fuel ratios (see
Chaps. 5 and 6).
The rationale underlying this infrastructure is that the tree wall will alter the
energy balance leading to a cooling within the firebreak via a large flux of latent
heat energy (evaporation). Decomposition rates are low due to the high humidity,
and litter is also kept at high moisture and with a high packing ratio, which further
hinders propagation.
Fire intensity should thus diminish after reaching a green firebreak. Some studies
report that green firebreaks are not only effective for controlling low surface fires.
For instance, the spread of a fire with 15-m-high flames was reported to be pre-
vented by a 10-m-wide firebreak of Schima superba (Cui et al. 2019).
Green firebreaks are particularly common in China. Given the physiological
requirements of the species within firebreaks, they cannot be placed anywhere. In
order to maintain the fast growth rates of the tree and the high evaporative require-
ments of this vegetation, this needs to be located under ample water supply. However,
their efficiency may be limited under spotting or plume-driven fires, as previously
mentioned when discussing linear infrastructures (Sect. 9.3).

9.6 Pyrophysiology: Plant Flammability and Fire Behavior

Understanding whether or not plant traits affect landscape patterns of fire behavior
has been the subject of intense discussions in the literature. Probably the two physi-
ological factors most important in affecting the activity of past wildfires have been
the percent of dead matter in the canopy and the degree of ladder fuels by self-­
pruning and by self-thinning (Chap. 6). These parameters are important as they
affect the probability and the spread of high-intensity canopy fires. Other factors,
such as mineral content, leaf moisture, or leaf morphological features, may play a
secondary role, particularly in affecting low-intensity surface fires.
However, the role of such physiological traits as drivers of fire activity may have
declined in recent years, as wildfires have become more virulent (Fig. 9.1). We have
previously explained the different fire generations as well as the different wildfire
types by fire spread patterns. The extreme fire behavior of, for instance,
9.6 Pyrophysiology: Plant Flammability and Fire Behavior 167

s­ ixth-­generation fires, as well as of plume-driven fires, is largely independent from


the species present in the landscape (Fernandes et al. 2019). These extreme fires
depend upon large fuel accumulations that release enough energy to interact with
tropospheric dynamics and alter wind currents, regardless of the species present.
Furthermore, the importance of fuel traits in other fire types, such as topographic
and wind-driven fires, is even lower than in convective fires. Studies assessing the
drivers of these fires note that they are largely driven by landscape patterns of land
cover and topography, rather than by fuel attributes (Duane et al. 2015).
That is not to say that ecophysiological processes do not play a role. Understanding
growth rates and predicting drought responses such as deposition as well as drought-­
induced mortality, for instance, are important to understand when critical thresholds
of fuel accumulation have been reached. By critical thresholds here we mean the
amount and structure of the fuel that, when burning, allow for the development of
massive plumes that lead to sixth-generation fires or plume-driven fires. Growth
rates and drought responses are crucial as they are major drivers of the fire energy
balance. However, we should note that additional factors, such as landscape patterns
of land cover, atmospheric dynamics, and topography, play a large role in under-
standing fire behavior, particularly once the critical fuel accumulation thresholds
have been reached.
There has been a large effort towards understanding the drivers of leaf-level
flammability, following the early definitions of flammability as composed of ignit-
ability, sustainability, combustibility, and consumability (see Chap. 4). Recent stud-
ies have expanded this approach by additionally incorporating variation under wind
tunnels and also by including other plant organs such as shoots or bark or whole
plants. This approach is valuable as it provides useful information at lower organi-
zational scales. However, one must make a considerable leap of faith in order to
believe that the mechanisms governing landscape patterns of wildfires are the same
as those occurring at the bench or in wind tunnels or that the scaling of these pro-
cesses across time and space will be linear.
The approach of experimenting at lower organization scales (e.g., leaves as
opposed to landscapes, which would be higher organization scales) is analogous to,
for instance, using greenhouses in plant ecophysiological research. Highly artificial
conditions such as growing plants in pots within a greenhouse may provide impor-
tant insights on the mechanisms underlying a pattern, and it also allows for the
isolation of different environmental factors. However, these studies are limited as
their results may be affected by factors such as pot size or the substrate used.
Furthermore, the energy balance is disrupted, and, consequently, the field applica-
tion of these studies is limited.
When it comes to understanding wildfire behavior, the limitations of these artifi-
cial approaches are much higher. Unlike plant growth, which is a plant phenome-
non, wildfires are a biosphere-atmosphere phenomenon. That is, wildfires depend
upon the interactions between land and also the atmosphere. Flammability at land-
scape scales does not only depend on plant traits but also on landscape topographic
patterns as well as soil and atmospheric conditions. We may encounter, for instance,
that the flammability of a given stand depends on environmental conditions at the
168 9 Pyrophysiology and Wildfire Management

time of the fire (Fig. 2.3) even when the community composition or structure did
not change.
Understanding how fire modifies tree structure, however, may help in under-
standing and reconstructing patterns of fire spread. There are two modifications over
tree structure that are particularly useful for this (Fig. 9.3). One is the charring
height, which is higher on the leeward side of the tree, and the accompanying trian-
gular fire scars that are sometimes formed at the base of the leeward side of the
trunk. As the fire encounters an obstacle, the flow becomes turbulent leading to
vortex formations at the leeward side increasing flame height and fire residence time
(Fig. 9.3a, b) (Gutsell and Johnson 1996). Fire scars can then be used for recon-
structions of the fire regime based on dendrochronological approaches and also for
inferring information on fire intensity.
The second effect of fire spread on tree structure that may help in reconstructing
fire activity is the pattern of branching after a fire. The radiant and convective heat
flux deforms the xylem structure such that branches are twigs point towards the
direction of the fire spread (Fig. 9.3c).
Summing up, landscape patterns of flammability are not solely a property of the
fuel, but a property of the heat transfer mechanisms within the fire environment.
Consequently, the effects of the fuel traits on landscape flammability are contingent
upon such fire environment. One must thus use caution when trying to extrapolate
the role of physiological traits over landscape fire patterns. Understanding how fires
modify tree structure may additionally serve to infer fire spread patterns and to
determine its point of origin.

Fig. 9.3 Understanding the effects of wildfire on tree structure is useful to reconstruct patterns of
fire spread. (a) Wildfires burn the stem to a higher height on the leeward side of the trunk (Photo
credit by J. Madrigal: http://fuegolab.blogspot.com/2017/04/el-efecto-chimenea-en-los-troncos-
de.html) and sometimes create a fire scar at the base (b). The branches bend following the direction
of fire spread (c)
9.7 Pyrophysiology and Fire Behavior Modeling 169

9.7 Pyrophysiology and Fire Behavior Modeling

The importance of pyrophysiology, in terms of the traits that we have mentioned in


this book, has been recognized for long within the fire behavior community (Van
Wagner 1977; Rothermel 1972), and it is also acknowledged in more recent models
(Linn et al. 2010; Mell et al. 2007). The traditional modeling approach assumes that
mostly fuel load and structure, rather than fuel traits, drive fire behavior in combina-
tion with wind, slope, and moisture. More recent approaches allow for explicit
incorporation of fire-atmosphere interactions (Mell et al. 2007; Linn et al. 2013),
and they coincide in the importance of fuel load and structure. The only possible
exception is the model of Zylstra et al. (2016), which proposed that traits are more
important drivers of fire behavior than fuel loads.
The degree of biological realism in the representation of fuel dynamics depends
on each model type. Some allow for very detailed representations of crown dynam-
ics (Parsons et al. 2011), but fuel is more often considered as a static parameter. Fire
modeling plays an essential component in the development of fire danger rating
systems as well as in the development of fuel management plans. Fire simulations
are often ran in order to determine the locations where fire prevention actions (e.g.,
SMAs) will be more critical as well as in understanding seasonal variation of
fire danger.
It should be noted that fire behavior models show, at the time of writing, notable
difficulties in recreating, particularly, the behavior of extremely large fires, such as
those including pyroconvective activity, which are the most damaging fire types
(Fig. 9.1), or spotting. It is unlikely that adding ecophysiological information into
these fire models may substantially improve model fit. As other authors have noted,
some of the major problems in fire behavior modeling lie in their empirical nature
as well as in the underlying assumptions on heat transfer (Finney et al. 2013; Finney
et al. 2015), but see Mell et al. (2007). Furthermore, the effect that plumes exert on
fire behavior and the interaction with atmospheric instability and the effects of pyro-
convection on fire behavior are far from being resolved. Increased ecophysiological
information is thus not a priority in current fire models.
Pyrophysiology however shows great potential in order to improve even further
the degree of biological realism in fuel dynamics and seasonal variation in fire risk.
Modeling the seasonal variations in key fuel traits, as well as growth, mortality, and
deposition rates, could be coupled with current fire behavior models to more fully
predict potential for temporal changes in fire dynamics and aid in forest planning
actions. Such models could, for instance, determine the years necessary for a thinned
stand to grow back to reach critical CBD thresholds, how drought is affecting the
proportion of dead to live fuels and plant mortality across seasons, the rates of seed-
ling growth after prescribed burns, or the potential effects of prescribed fires.
The challenge lies, on the one hand, on merging physiological models, which are
often focused at tree or stand scales, with fire models, which require landscape or
even regional forecasts. On the other hand, there are some processes over which
plant ecophysiology has not yet paid much attention. For instance, the drivers of leaf
170 9 Pyrophysiology and Wildfire Management

senescence, particularly pre-programmed shedding in the fire season (Fig. 5.3), is


not properly understood, and, consequently, this limits model development.

9.8  yrophysiology and Post-fire Management


P
and Restoration

Ecophysiology and restoration are two disciplines that have traditionally developed
independently from each other (Moreira et al. 2012), although an increasing number
of studies have appeared in the last few years. Furthermore, ecophysiology is of
great value to understand whether and how restoration actions are necessary as well
as to evaluate the success of these restoration efforts.
One of the earliest management decisions that need to be made after a fire is
whether the remaining stand should be logged or not. There are different elements
that need to be considered before making such decision, and it would be beyond
scope to enter that discussion here. However, ecophysiological measurements can
help us understand, from the point of view of the regenerating seedlings, which is
the best strategy (Moya et al. 2015).
An example is the study by Marañón-Jiménez et al. (2013) which compared
naturally regenerating seedlings of P. pinaster under three treatments: no interven-
tion; logging and pruning trees with biomass left on site without mastication; and
logging and masticating the debris (salvage logging). They observed that seedlings
in the intermediate treatment, where trees were cut but not masticated, showed the
highest growth rates and pine cone numbers 6 years after the fire. They examined
the stable isotope composition of oxygen in leaf bulk tissue, an indicator of water
use, and observed a higher stomatal conductance in that treatment, indicating that it
had the highest water availability. They also examined the stable isotope composi-
tion of carbon in leaf bulk tissue along with leaf nitrogen, indicators of photosyn-
thetic carbon gain, and also observed a higher photosynthesis in the intermediate
treatment. Finally, they examined leaf nutrition and observed that the growth
improvement in the intermediate treatment was only driven by the amelioration in
the water and energy balances, as the nutrient status was not different across
treatments.
When planting tree species for restoration is necessary, there are a number of
factors to take into account. The most important aspect will be climate change. As
temperatures raise due to climate warming, we might then need to select for species,
or for provenances, that are currently occurring at warmer or drier locations (Resco
de Dios et al. 2007), a process that has sometimes been termed as assisted migration
(Mclachlan et al. 2007). This contrasts with the traditional approach where species
were introduced during planting efforts based upon their suitability to thrive under
past climate conditions.
In fact, it is common to encounter species that, to a major or lesser degree, show
local adaptation to current conditions. As climate changes, we can expect a m­ ismatch
9.8 Pyrophysiology and Post-fire Management and Restoration 171

between species locally adapted to current climate and future climatic conditions.
Actions to solve such mismatch require understanding the physiological responses
to future climates for different species and from different provenances. Actions such
as ecosystem engineering or assisted migration will thus need to be based on an
understanding of the physiological responses to new conditions.
Some authors propose alternative approaches to solve this mismatch by changing
the ectomycorrhizal community (Gehring et al. 2017). That is, using inoculum from
dry-resistant mother trees may expand the regeneration niche of that species by
providing additional drought resistance. However, evidence for this process is still
in its infancy, and the underlying mechanisms still need to be resolved.
Additionally, one should also take into account the degree of genetic variation in
traits related to species survival and reproduction when selecting provenances (see
Chap. 4). There is variation in the degree of serotiny, for instance, and post-fire
recruitment in restored areas will be very different depending on this trait (Gil et al.
2009). Similarly, genetic variations in bark thickness (Madrigal et al. 2019), hydrau-
lic architecture, resprouting vigor, and other factors affecting persistence may also
be used to our favor in restoration activities.
Ecophysiology is also useful when attempting to evaluate the evolution of eco-
system function after fire. Current approaches to evaluate restoration efforts are
based on comparing compositional changes before and at different times after the
perturbation (Sánchez-Pinillos et al. 2019; Alday and Marrs 2014). Additionally
monitoring carbon assimilation and water use, for instance, we can evaluate to
which extent are key ecosystem properties such as net ecosystem exchange, evapo-
transpiration, or rain use efficiency recovering after fire. Ecophysiology thus may
take current evaluation efforts one step further.
Along these lines, ecophysiology may also shed light on the processes explain-
ing why different restoration techniques work or fail. Major efforts towards under-
standing the roles of facilitation or mulches have been made, to name a couple. In
some cases, but not all, restoration success is enhanced by the use of nurse shrubs
(Gómez-Aparicio et al. 2005). Concomitantly, a substantial body of research has
arisen to better understand how mulches and different soil conditioning techniques
improve restoration success (Coello et al. 2018). Evaluating the success or failure of
these practices would greatly benefit from incorporating physiological measure-
ments to understand the underlying processes.
Ecophysiology is of great importance in nurseries, for instance, as plants are
often exposed to some degree of stress for “hardening.” That is, subsequent drought
performance is ameliorated after an initial exposure to drought, a common practice
in nurseries. However, the intensity of drought necessary to trigger hardening varies
markedly across species. A drought too intense may kill the seedling, and one too
soft may not induce hardening. Understanding the physiological traits of each spe-
cies is thus essential for such hardening practices.
Ecophysiology may also be used in restoration efforts that seek to diminish the
degree of plant invasion after fire. The hypotheses of limiting similarity and the
principle of competitive exclusion propose that two species occupying the same
niche cannot coexist indefinitely. Based upon this hypothesis, some restoration
172 9 Pyrophysiology and Wildfire Management

efforts have been attempted where different trait combinations are introduced in
efforts to reduce the potential success of plant invaders. This can be relevant in areas
where invasives may be favored by fire. Cheatgrass (Bromus tectorum) invasion in
the Great Basin, USA, for instance, enhances fire activity by creating a dry layer of
grass at the end of the summer which is readily consumed by fire. Native species, in
contrast, lack fire adaptations to survive under highly frequent fires. The result of
this process is that native species are being replaced by cheatgrass modifications on
the fire cycle. One possible way to control this invasion would be the post-fire intro-
duction of trait combinations that hinder cheatgrass establishment. Although there
have been some attempts that control invasions in this manner, it should be men-
tioned that evidence on its efficiency is still scarce (Kimball et al. 2016).
The examples on how ecophysiology may aid in post-fire management are
numerous, too many to be cited here. Whereas ecophysiology is not often consid-
ered as relevant for post-fire restoration, this list of examples will hopefully serve to
exemplify the great potential in merging both knowledge areas.

References

Agee JK, Skinner CN (2005) Basic principles of forest fuel reduction treatments. For Ecol Manag
211:83–96. https://doi.org/10.1016/j.foreco.2005.01.034
Alcasena FJ, Ager AA, Salis M, Day MA, Vega-Garcia C (2017) Optimizing prescribed fire allo-
cation for managing fire risk in central Catalonia. Sci Total Environ 621:872–885. https://doi.
org/10.1016/j.scitotenv.2017.11.297
Alday JG, Marrs RH (2014) A simple test for alternative states in ecological restoration: the use
of principal response curves. Appl Veg Sci 17(2):302–311. https://doi.org/10.1111/avsc.12054
Battipaglia G, Savi T, Ascoli D, Castagneri D, Esposito A, Mayr S, Nardini A (2016) Effects of
prescribed burning on ecophysiological, anatomical and stem hydraulic properties in Pinus
pinea L. Tree Physiol. https://doi.org/10.1093/treephys/tpw034
Boer MM, Sadler RJ, Wittkuhn RS, McCaw L, Grierson P (2009) Long-term impacts of prescribed
burning on regional extent and incidence of wildfires—evidence from 50 years of active fire
management in SW Australian forests. For Ecol Manag 259:132–142
Boer MM, Price OF, Bradstock R (2015) Wildfires: weigh policy effectiveness. Science 350:920
Brennan TJ, Keeley JE (2017) Impacts of mastication fuel treatments on California, USA, chap-
arral vegetation structure and composition. Fire Ecol 13(2):120–138. https://doi.org/10.4996/
fireecology.130312013
Castellnou M, Miralles M (2010) The changing face of wildfires. Crisis Resp 5(56):57
Castellnou M, Guiomar N, Rego F, Fernandes PM (2018) Fire growth patterns in the 2017
mega fire episode of October 15, central Portugal. In: Viegas DX (ed) Advances in forest fire
Research 2018. Imprensa da Universidade de Coimbra, Coimbra, pp 447–453. https://doi.
org/10.14195/978-989-26-16-506_48
Coello J, Ameztegui A, Rovira P, Fuentes C, Piqué M (2018) Innovative soil conditioners and
mulches for forest restoration in semiarid conditions in northeast Spain. Ecol Eng 118:52–65.
https://doi.org/10.1016/j.ecoleng.2018.04.015
Costa P, Castellnou M, Larrañaga A, Miralles M, Kraus PD (2011) Prevention of large wildfires
using the fire types concept. Unitat Tècnica del GRAF, Cerdanyola del Vallès, Barcelona
Cui X, Alam MA, Perry GL, Paterson AM, Wyse SV, Curran TJ (2019) Green firebreaks as a
management tool for wildfires: lessons from China. J Environ Manag 233:329–336. https://doi.
org/10.1016/j.jenvman.2018.12.043
References 173

Duane A, Piqué M, Castellnou M, Brotons L (2015) Predictive modelling of fire occurrences from
different fire spread patterns in Mediterranean landscapes. Int J Wildland Fire 24:407–418.
https://doi.org/10.1071/wf14040
Fernandes PM, Guiomar N, Rossa CG (2019) Analysing eucalypt expansion in Portugal as a fire-­
regime modifier. Sci Total Environ 666:79–88. https://doi.org/10.1016/j.scitotenv.2019.02.237
Fernandes PM, Delogu GM, Leone V, Ascoli D (2020) 10 – Wildfire policies contribution to foster
extreme wildfires. In: Tedim F, Leone V, McGee TK (eds) Extreme wildfire events and disas-
ters. Elsevier, pp 187–200. https://doi.org/10.1016/B978-0-12-815721-3.00010-2
Fernández C, Fernández-Alonso JM, Vega JA (2019) Effects of mastication of burned non-­
commercial Pinus pinaster Ait. trees on soil compaction and vegetation response. For Ecol
Manag 449:117457. https://doi.org/10.1016/j.foreco.2019.117457
Finney MA, Cohen JD, McAllister SS, Jolly WM (2013) On the need for a theory of wildland fire
spread. Int J Wildland Fire 22(1):25–36. https://doi.org/10.1071/wf11117
Finney MA, Cohen JD, Forthofer JM, McAllister SS, Gollner MJ, Gorham DJ, Saito K, Akafuah
NK, Adam BA, English JD (2015) Role of buoyant flame dynamics in wildfire spread. Proc
Natl Acad Sci 112(32):9833–9838. https://doi.org/10.1073/pnas.1504498112
Gehring CA, Sthultz CM, Flores-Rentería L, Whipple AV, Whitham TG (2017) Tree genetics
defines fungal partner communities that may confer drought tolerance. Proc Natl Acad Sci
114(42):11169–11174. https://doi.org/10.1073/pnas.1704022114
Gil L, López R, García-Mateos Á, González-Doncel I (2009) Seed provenance and fire-related
reproductive traits of Pinus pinaster in central Spain. Int J Wildland Fire 18(8):1003–1009.
https://doi.org/10.1071/WF08101
Gómez-Aparicio L, Valladares F, Zamora R, Quero JL (2005) Response of tree seedlings to the
abiotic heterogeneity generated by nurse shrubs: an experimental approach at different scales.
Ecography 28(6):757–768
Gutsell S, Johnson E (1996) How fire scars are formed: coupling a disturbance process to its eco-
logical effect. Canadian journal of Forest research-revue Canadienne De Recherche Forestiere.
Can J For Res 26:166–174. https://doi.org/10.1139/x26-020
Hood S, Sala A, Heyerdahl EK, Boutin M (2015) Low-severity fire increases tree defense against
bark beetle attacks. Ecology 96:1846–1855
Hood SM, Baker S, Sala A (2016) Fortifying the forest: thinning and burning increase resistance
to a bark beetle outbreak and promote forest resilience. Ecol Appl 26(7):1984–2000. https://
doi.org/10.1002/eap.1363
IPCC (2014) Summary for policymakers. In: Field CB, Barros VR, Dokken DJ et al. (eds) Climate
change 2014: impacts, adaptation, and vulnerability. Part A: global and sectoral aspects.
Contribution of Working Group II to the fifth assessment report of the intergovernmental panel
on climate change, Cambridge University Press, Cambridge, pp 1–32
Jiménez E, Vega-Nieva D, Rey E, Fernández C, Vega JA (2016) Midterm fuel structure recovery
and potential fire behaviour in a Pinus pinaster Ait. forest in northern central Spain after thin-
ning and mastication. Eur J For Res 135:675–686. https://doi.org/10.1007/s10342-016-0963-x
Kimball S, Funk JL, Sandquist DR, Ehleringer JR (2016) Ecophysiological considerations for res-
toration. In: Palmer MA, Zedler JB, DA Falk (eds) Foundations of restoration ecology, Island
Press, Washington, pp 153–181. https://doi.org/10.5822/978-1-61091-698-1_6
Linn RR, Winterkamp JL, Weise DR, Edminster C (2010) A numerical study of slope and fuel
structure effects on coupled wildfire behaviour. Int J Wildland Fire 19:179–201
Linn RR, Sieg CH, Hoffman CM, Winterkamp JL, McMillin JD (2013) Modeling wind fields
and fire propagation following bark beetle outbreaks in spatially-heterogeneous pinyon-­
juniper woodland fuel complexes. Agric For Meteorol 173:139–153. https://doi.org/10.1016/j.
agrformet.2012.11.007
Madrigal J, Souto-García J, Calama R, Guijarro M, Picos J, Hernando C (2019) Resistance of
Pinus pinea L. bark to fire. Int J Wildland Fire. https://doi.org/10.1071/wf18118
Marañón-Jiménez S, Castro J, Querejeta JI, Fernández-Ondoño E, Allen CD (2013) Post-­
fire wood management alters water stress, growth, and performance of pine regeneration
174 9 Pyrophysiology and Wildfire Management

in a Mediterranean ecosystem. For Ecol Manag 308:231–239. https://doi.org/10.1016/j.


foreco.2013.07.009
Mclachlan JS, Hellmann JJ, Schwartz MW (2007) A framework for debate of assisted
migration in an era of climate change. Conserv Biol 21(2):297–302. https://doi.
org/10.1111/j.1523-1739.2007.00676.x
Mell W, Jenkins MA, Gould J, Cheney P (2007) A physics-based approach to modelling grassland
fires. Int J Wildland Fire 16(1):1–22. https://doi.org/10.1071/WF06002
Molina-Terrén DM, Xanthopoulos G, Diakakis M, Ribeiro L, Caballero D, Delogu GM, Viegas
DX, Silva CA, Cardil A (2019) Analysis of forest fire fatalities in southern Europe: Spain,
Portugal, Greece and Sardinia (Italy). Int J Wildland Fire 28(2):85. https://doi.org/10.1071/
wf18004
Moreira F, Arianoutsou M, Corona P, Heras JDl (2012) Post-fire management and restoration of
Southern European forests, vol 24. Managing Forest ecosystems. Springer, Cham
Moreno-Gutierrez C, Barbera GG, Nicolas E, M DEL, Castillo VM, Martinez-Fernandez
F, Querejeta JI (2011) Leaf delta18O of remaining trees is affected by thinning inten-
sity in a semiarid pine forest. Plant Cell Environ 34(6):1009–1019. https://doi.
org/10.1111/j.1365-3040.2011.02300.x
Moritz MA, Batllori E, Bradstock RA, Gill AM, Handmer J, Hessburg PF, Leonard J, McCaffrey
S, Odion DC, Schoennagel T, Syphard AD (2014) Learning to coexist with wildfire. Nature
515:58–66. https://doi.org/10.1038/nature13946
Moya D, Heras J, López-Serrano F, Ferrandis P (2015) Post-fire seedling recruitment and morpho-­
ecophysiological responses to induced drought and salvage logging in Pinus halepensis Mill.
stands. Forests 6(12):1858–1877. https://doi.org/10.3390/f6061858
Parsons RA, Mell WE, McCauley P (2011) Linking 3D spatial models of fuels and fire: effects of
spatial heterogeneity on fire behavior. Ecol Model 222(3):679–691. https://doi.org/10.1016/j.
ecolmodel.2010.10.023
Perrakis DDB, Agee JK, Eglitis A (2011) Effects of prescribed burning on mortality and resin
defenses in old growth ponderosa pine (Crater Lake, Oregon): four years of post-fire monitor-
ing. Nat Areas J 31:14–25
Piqué M, Coello J, Amari Y, Aletà N, Sghaier T, Mutke S (2017) Grafted stone pine plantations for
cone production: trials on Pinus pinea and Pinus halepensis rootstocks from Tunisia and Spain.
Options Méditérr 122:17–23
Price OF, Penman TD, Bradstock RA, Boer MM, Clarke H (2015) Biogeographical variation in the
potential effectiveness of prescribed fire in south-eastern Australia. J Biogeogr 42(11):2234–
2245. https://doi.org/10.1111/jbi.12579
Resco de Dios V, Fischer C, Colinas C (2007) Climate change effects on Mediterranean forests and
preventive measures. New For 33:29–40
Ribeiro LM, Viegas DX, Almeida M, McGee TK, Pereira MG, Parente J, Xanthopoulos G, Leone V,
Delogu GM, Hardin H (2020) 2 – Extreme wildfires and disasters around the world: lessons to be
learned. In: Tedim F, Leone V, McGee TK (eds) Extreme wildfire events and disasters. Elsevier,
Amsterdam/Cambridge, MA, pp 31–51. https://doi.org/10.1016/B978-0-12-815721-3.00002-3
Rodríguez-García A, Madrigal J, González-Sancho D, Gil L, Guijarro M, Hernando C (2018)
Can prescribed burning improve resin yield in a tapped Pinus pinaster stand? Ind Crop Prod
124:91–98. https://doi.org/10.1016/j.indcrop.2018.07.049
Rothermel RC (1972) A mathematical model for predicting fire spread in wildland fuels. USDA
Forest Service, Intermountain Forest and Range Experiment Station, Research Paper INT-115.
Ogden, UT
Sala A, Peters GD, McIntyre LR, Harrington MG (2005) Physiological responses of ponderosa pine
in western Montana to thinning, prescribed fire and burning season. Tree Physiol 265:329–348
Sánchez-Pinillos M, Leduc A, Ameztegui A, Kneeshaw D, Lloret F, Coll L (2019) Resistance,
resilience or change: post-disturbance dynamics of boreal forests after insect outbreaks.
Ecosystems. https://doi.org/10.1007/s10021-019-00378-6
Soler Martin M, Bonet JA, Martínez De Aragón J, Voltas J, Coll L, Resco De Dios V (2017)
Crown bulk density and fuel moisture dynamics in Pinus pinaster stands are neither modified
References 175

by t­hinning nor captured by the Forest Fire Weather Index. Ann For Sci 74:51. https://doi.
org/10.1007/s13595-017-0650-1
Sparks AM, Smith AMS, Talhelm AF, Kolden CA, Yedinak KM, Johnson DM (2017) Impacts of
fire radiative flux on mature Pinus ponderosa growth and vulnerability to secondary mortality
agents. Int J Wildland Fire 26(1):95. https://doi.org/10.1071/wf16139
Valor T, Casals P, Altieri S, González-Olabarria JR, Piqué M, Battipaglia G (2018) Disentangling
the effects of crown scorch and competition release on the physiological and growth response
of Pinus halepensis Mill. using δ 13 C and δ 18 O isotopes. For Ecol Manag 424:276–287.
https://doi.org/10.1016/j.foreco.2018.04.056
Van Wagner CE (1977) Conditions for the start and spread of a crown fire. Can J For Res 7:23–34
Zausen GL, Kolb TE, Bailey JD, Wagner MR (2005) Long-term impacts of stand management on
ponderosa pine physiology and bark beetle abundance in northern Arizona: a replicated land-
scape study. For Ecol Manag 218:291–305
Zylstra P, Bradstock RA, Bedward M, Penman TD, Doherty MD, Weber RO, Gill AM, Cary GJ
(2016) Biophysical mechanistic modelling quantifies the effects of plant traits on fire severity:
species, not surface fuel loads, determine flame dimensions in eucalypt forests. PLoS One
11(8):e0160715. https://doi.org/10.1371/journal.pone.0160715
Chapter 10
Global Change, Pyrophysiology,
and Wildfires

Abstract We can expect many changes in the activity of wildfires as a result of


global change. Most research has focused on climatic effects, but additional factors
such as changes in land cover and land use and including plant biological invasions
also play major roles. Furthermore, wildfires exert a large influence to affect global
biodiversity, and, depending on management, the effects may be either positive or
negative. Here we begin by explaining geographical variations in fire activity as a
function of first biogeographical principles and which changes in fire activity we
might expect globally. We will then discuss more specifically the issues associated
with each of the major biomes globally. Next, we will move to discussing the new
challenges for civil protection that fires will bring. We will pay particular attention
to issues related to the wildland urban interface and to pyroconvective activity. The
chapter ends by discussing the interactions between wildfire activity and invasive
species and biodiversity.

10.1 Fire and Climate Change

In order to understand how climate change is going to affect the fire cycle, we need
to develop models capable of projecting biogeographic variations in fire regimes.
That is, we need to develop robust models that accurately capture the biophysical
principles explaining global dynamics of fire activity (Boer et al. 2019; Bradstock
2010). Fire activity varies across productivity gradients, such that high moisture
limits fire activity in wet forests and fuel activity limits fire activity at xeric sites
(Fig. 1.3) (Krawchuk and Moritz 2011; Boer et al. 2016). We can thus consider that
fire activity across biomes may be limited by either fuel productivity (PL) or fuel
dryness (DL) (Fig. 10.1) (Boer et al. 2016; Boer et al. 2019). Productivity-limited
type fire ecosystems are those where fuel loads are generally low and, consequently,
fires occur after wet seasons that foster fuel build-up, that is, grassy and, potentially,
shrubby biomes. DL-type fire ecosystems, in contrast, are those with high fuel accu-
mulations and where such fuel is usually not available to the fire because of its high
moisture, that is, forests and, specially, those with ladder fuels.

© Springer Nature Switzerland AG 2020 177


V. Resco de Dios, Plant-Fire Interactions, Managing Forest Ecosystems 36,
https://doi.org/10.1007/978-3-030-41192-3_10
178 10 Global Change, Pyrophysiology, and Wildfires

Fig. 10.1 Geographical distribution of domains of productivity (PL)- and dryness (DL)-type fire
in green and orange tones, respectively. In white land areas, the observed mean annual fractional
burned area was negligible (F < 0.000001). (Reproduced from Boer et al. (2019))

Fig. 10.2 Expected changes in fire activity based on a multimodel ensemble for the end of the
century. (Reproduced from Moritz et al. (2012))

There are three types of changes in fire activity that we may encounter with climate
and global change: changes in the type of fire, in area burned, and in fire ­intensity.
Regarding the first change, transitions from DL- to PL-fire type may occur in areas cur-
rently occupied by forests or shrublands, where future aridity may limit fuel produc-
tion. In some cases a shift from PL- to DL- type fires might also occur (Boer et al. 2016).
Changes in fire activity are dependent on the fire type. For instance, we can
expect that fire activity will increase with aridity in DL-type fire ecosystems, but not
in Productivity-limited type fire (Fig. 10.2). Overall we can expect a poleward and
upward migration of the fire-belt in Europe. That is, while fire activity in Europe is
currently limited to the Mediterranean arc, we can expect an increase in fire activity
10.1 Fire and Climate Change 179

in Northern countries as a result of increasing drought. Additionally, whereas fire


was traditionally rare in mountain regions such as the Pyrenees or the Alps, we can
expect increases in fire activity in these areas if fuel becomes increasingly available.
Consistent with this hypothesis, we have started to observe in recent years the
occurrence of wildfires in both northern countries as well as in mountainous regions
(Cardil et al. 2016). Similarly, we might expect increases in fire activity also in
American conifer forests as well as in Boreal forests, when aridity increases, as they
are limited by fuel availability (Fig. 10.2) (Moritz et al. 2012). In Australia, increases
in fire activity in forest areas such as those occurring the southeast may lead to
transforming those ecosystems from DL- to Productivity-limited type fires (Boer
et al. 2016). Changes in fire activity in tropical and subtropical forests are more
linked to changes in land use or land cover, as we will discuss later in this chapter.
In addition to changes in fire frequency and in fire type, we might also experience
increases in fire intensity, potentially leading to increases in catastrophic mega-fires,
in DL-type fire ecosystems that currently show limited fire activity as they are too
wet to burn. This is the case of temperate forests (and potentially tropical forests
too, but here other factors affect the response as will be discussed in the next sec-
tion), which could experience pyroconvection activity, and an increasing frequency
in sixth-generation fires (Chap. 9). This is because of the high fuel loads available
in these ecosystems. There are some reports already of increased fire activity in
temperate forests as a result of climate change (Westerling et al. 2006), and those
could intensify in areas with enough fuel loads.
Overall, we may expect wildfires to fasten the rate at which terrestrial ecosys-
tems respond to drought-induced climate change. In the absence of disturbance,
many forests and shrublands decline and are replaced by more drought-resistant
communities. In fact, the number of reports of drought-induced forest mortality has
been increasing over time (Allen et al. 2015). Wildfires may facilitate such transi-
tion, but they could also lead to increased deforestation in areas where there is no
propagule capable of regenerating under fire or under the novel environmental con-
ditions created by climate change (Davis et al. 2019; Karavani et al. 2018). We now
discuss, in more detail, projected changes to fire activity across biomes.

10.1.1  ire in Tropical Forests Is Linked to Land Cover


F
and Land Use

Wildfires are not native to tropical forests, and their presence is due to human activ-
ity. Indeed, current projections based on climate and fuels project further decreases
in burned area (Fig. 10.2). Wildfires have been mostly absent from the geological
record in tropical forests, and they show very low return intervals, occurring once
every 400–1000 years, at most. Consequently tropical trees lack adaptations to fire.
They show very thin barks that do not protect the buds from fire-induced thermal
stress (Charles-Dominique et al. 2017; Rosell 2016). Tropical wildfires often burn
very slowly, advancing a few hundred meters per day. They burn at very low inten-
180 10 Global Change, Pyrophysiology, and Wildfires

sity, with flames that only seldom pass above the knee. However, such low-intensity
fires show very high severity, as residence times are high and also because of above-
mentioned lack of adaptations to survive fire (Cochrane and Schulze 1998).
Consequently, fires in tropical forests lead to a change in land cover and, for the case
of the Amazon, to a savannization (Nobre et al. 2016). Furthermore, changes in
vegetation structure after the first fire lead to an increase in pioneer species and to
changes in composition that increase the intensity of subsequent fires (Cochrane
et al. 1999).
Fires in the Amazon are not a natural phenomenon. They are used as a tool for
land clearing and subsequent planting of soy or other crops. However, the previ-
ously mentioned projections of declines in fire activity (Fig. 10.2) may be com-
pletely overridden by human-driven ignitions associated with deforestation. The
future of wildfires in the Amazon is thus driven to a larger extent by human practices
and land use policies than by climate, fuel, or other factors (Barlow et al. 2019).
Wildfires in SE Asia also result from human-induced land use change. But in this
case they related to peat draining (Turetsky et al. 2015). Peat fires dominate in coun-
tries like Indonesia in Malaysia, and they occur after land drainage for planting
palm oil and other crops. These areas are prone to peat fires given the vast accumu-
lations of belowground organic matter that, as drought increases, become suitable
for wildfires. Climate change and increased aridity further increase fuel availability
(Turetsky et al. 2015). There have been some governmental measures to restore
degraded peatlands, and future fire regimes will depend upon governmental mea-
sures to recover groundwater levels and native vegetation. In the absence of such
measures, climate change will increase fire activity.
Tropical forests from the Congo Basin or South Asia are, at the time of writing,
not threatened by human activities. Deforestation rates in the Congo Basin are about
half of those in the Amazon forests (FAO 2011). However, such forests are undergo-
ing a drought-induced change in species composition (Zhou et al. 2014). This effect
is not captured by current models, and future assessments would need to address the
carry-over effects of such species replacements and aridification on the fire regime.
In tropical and subtropical China, wildfires are mostly confined to conifer and mixed
forests that experience seasonal drought (Ying et al. 2018). The main challenge here
is, similar to that in the Congo Basin, the potential consequences of changes in spe-
cies composition as a result of climate change.

10.1.2  oody Thickening Affects Tropical and Subtropical


W
Savanna Fires

Savannas concentrate 70% of the area burned globally (Giglio et al. 2013). Any fac-
tor altering the structure of the vegetation in savannas, which feeds back to affect its
fire cycle, may exert large carry-over effects on global fire activity. Woody
10.1 Fire and Climate Change 181

t­ hickening, the encroachment of woody plants in grassy areas or the increase in the
density of woody plants in woodlands and savannas, could be such a factor.
High-frequency savanna fires rely on rainy seasons that produce enough grassy
fuels to carry a fire. As the percentage of woody cover increases, we may expect a
decline in fire activity and a shift in the system from PL- to DL-type fire.
Woody thickening occurs in all continents with tropical and subtropical savan-
nas. Reported rates of annual increment in woody cover range from 0.1% in
Australian savannas to 0.3% in African savannas and 1% in South American savan-
nas (Stevens et al. 2016).
The question is thus what is driving such woody thickening and whether it will
continue into the future. Different hypotheses have been raised to answer this ques-
tion, and they include increased CO2 concentrations, altered rainfall regimes, biotic
facilitations, and changes in the fire regime (Van Auken 2009).
Given the global occurrence of woody thickening, some authors have proposed
that a global mechanism, such as raising CO2 concentrations, is driving the response.
There are two potential reasons why could CO2 increase woody cover. One is that
stomatal conductance often declines in response to CO2, and this could lead to a
decline in transpiration which could potentially increase the water in deeper soil
areas, where only woody roots have access (Morgan et al. 2007).
The second mechanism is related to the effects of CO2 on plant growth and
reserves. Mesic savannas may burn as frequently as every 1–3 years. Such high fire
frequency creates a trap for seedlings that are topkilled by fire. Bond and Midgley
(2000) proposed that increasing CO2 concentrations could allow for faster growth
rates such that trees would be capable of reaching earlier the critical height. That is,
they proposed that trees taller than 3 m avoid topkill by fire. If elevated CO2 increases
plant growth, because it fosters photosynthesis, we could expect saplings reaching
the 3 m threshold faster. Additionally, elevated CO2 could increase carbohydrate
reserves and post-fire resprouting would not be limited by carbohydrates.
Free-Air CO2 Enrichment (FACE) facilities, where intact ecosystems are
exposed to elevated CO2 concentrations, provide direct opportunities to test this
hypothesis. These experiments show, on the one hand, that water savings some-
times, but not always, occur as other adjustments beyond stomatal behavior are
responsible for tree transpiration and the soil water balance (Gimeno et al. 2018).
On the other hand, CO2 effects over growth are limited to areas with high nutrient
availability. When fertility limits growth, however, potential CO2 benefits are lim-
ited, and they may not affect seedling growth or carbohydrate reserves. In a Free-­
Air CO2 Enrichment Experiment under a putatively phosphorus-limited soil, Collins
et al. (2018) observed that sapling growth of Eucalyptus tereticornis and Hakea
sericea was more affected by competition and herbivory than by CO2. Additionally,
they also observed that resprouting was linked to lignotuber volume, but that was
not a function of CO2.
Another factor that could lead to woody plant encroachment is interactions
between herbivore frequency and the fire cycle. This factor is thought to play a role,
particularly in subtropical grasslands from SW USA (Van Auken 2009; Archer et al.
2017). Cattle introduction by European settlements at the turn of the nineteenth
182 10 Global Change, Pyrophysiology, and Wildfires

century, which prefer grasses over woody plants, would have broken the spatial
continuity of grassy fuel, thereby reducing fire frequency and favoring woody plant
establishment. Similar trends have been reported in Africa where increasing the
density of grazers correlated with highest rates of woody plant encroachment at a
continental scale analysis (Venter et al. 2018). These authors also reported that
woody plant encroachment may decline at the highest densities of grazers, presum-
ably because of trampling and other damages on woody plants.
Browsers (e.g., elephants), on the other hand, show the opposite effect, and
woody plant encroachment correlated negatively with browser densities. This is
because of direct effects on tree mortality or by reducing tree height and increasing
the probability of topkill by fire (Venter et al. 2018). In addition to changing species
composition, elephant paths may act as firebreaks by creating discontinuities in the
fuel, which would favor woody plant encroachment.
The notion that woody plant encroachment is driven by interactions between
herbivores and fire, however, has also not always been supported by experimental
evidence. Some studies conducted in SW USA and in S Africa report that rates of
woody plant encroachment inside of fenced areas, where grazers and browsers are
kept out, are no different than those outside of fenced areas (Bond and Midgley
2012; McClaran et al. 2003).
Additional factors come into play to alter rates of woody plant encroachment.
Processes such as landscape geomorphology and soil texture exert dramatic effects
over the water balance, which might impact woody plant recruitment. In the Sonoran
desert, for instance, higher rates of woody plant establishment are observed over
sandy than over loamy soils (Resco de Dios et al. 2012). Active and passive grass-­
tree interactions create additional landscape heterogeneity. Active grass-tree inter-
actions often lead to grasses limiting, but not inhibiting, woody seedling recruitment.
Passive interactions, where grasses die and become litter, may facilitate woody
plant establishment by limiting light and water competition with seedlings (Resco
de Dios et al. 2014).
These different hypotheses should not be viewed as antagonistic as different pro-
cesses may prevail over different areas. For instance, Bond and Midgley (2012)
propose that CO2 may be a more important driver of woody encroachment over
mesic savannas, whereas the effect of grazers may prevail on arid savannas.
Understanding the mechanisms underlying woody plant encroachment is not a
fully resolved problem, which hinders projections on its future evolution and, con-
sequently, may bring more uncertainty on the future of the fire regime in this area.
What seems to be clear is that fire has been an important element in maintaining
savannas in some areas. Fire as control of woody vegetation may lose its effective-
ness if woody plant encroachment continues, which would lead to a negative trend
of burned area over time.
10.1 Fire and Climate Change 183

10.1.3  nimodal Evolution of Wildfires in Mediterranean


U
Biomes

Mediterranean ecosystems will likely experience a unimodal pattern of fire activity


across the twenty-first century due to interactions with future climate and plant
physiological drought responses (Fig. 10.3) (Batllori et al. 2013). In the short term,
we can expect increases in fire activity if forest area continues to increase as a result
of land abandonment and if global warming leads to lower water availability over a
longer part of the year. This will consequently increase the total amount of canopy
fuels, as well as the proportion of dead canopy fuels. We might then observe a dry-
ing down of the vegetation, which increases fuel availability as a result of drought-­
induced senescence and, over longer periods of time, that might lead to tree and
forest mortality (Fig. 10.3). A number of reports already indicate an increased fre-
quency of climate-induced forest mortality, which might temporarily increase fuel
availability (Chap. 6).
Fire risk might decline over longer-term periods if climate change intensifies as we
would shift from a DL- to a Productivity-limited type fire. That is, some projections
indicate substantial reductions in vegetation cover across many Mediterranean regions,
including an increased likelihood of desertification in the Mediterranean basin

Forest mortality
Fire activity

 Dead:live
Less fuel loads

Cavitation & With fires Smaller plants


Shedding
Drought deciduous
Lower FMC Sparser vegetation
Species replacements
Increased
biomass
Insect attacks
Widespread defoliation

Time (intensification of climate change)


DL-fire type PL-fire type
Fig. 10.3 Projected changes in fire regime for Mediterranean ecosystems. Over time we can
expect fire activity in Mediterranean ecosystems to switch from being limited by fuel moisture
(DL-type) to being limited by fuel production (PL-type). The shape of temporal changes in fire
activity may be affected by drought and species replacements, and it may be more intense under
insect attacks and widespread defoliation. (Modified from Curt et al. (2013))
184 10 Global Change, Pyrophysiology, and Wildfires

(Cramer et al. 2018). If vegetation cover becomes more sparse, we may then expect a
decline in fire activity.
Understanding the transient response of vegetation to increasing drought is
remarkably complex as a myriad of factors may alter the response. We might
observe, for instance, species substitutions. In some areas of NE Spain, for instance,
a replacement of holm oak (Q. ilex) forests by Phillyrea latifolia shrublands is
underway (Peñuelas et al. 2017). These transitions might alter the availability of
canopy and surface fuels and exert additional feedbacks on the fire regime.
Mediterranean areas have also experienced an increasing defoliating trend over
the past 30 years or more (Carnicer et al. 2011). Increased defoliation may be the
result of drought-induced deciduousness and other processes discussed in Chap. 6,
and over long time scales it would lead to a decline in fire activity, due to the reduc-
tions in fuel load. It is worth noting that, under some circumstances, the increase in
dead biomass could also lead to either mega-fires, due to the large amounts of dead
biomass (Stephens et al. 2018), but lack of canopy fuels render this scenario less
likely (but still possible).
As drought stress intensifies, we might also expect interactive effects from pests
and pathogens that could further decrease fire activity. For instance, increases in the
presence of the defoliating pine processionary moth (Thaumetopoea pityocampa),
as documented in many areas of the Mediterranean (Hódar et al. 2003), might
decrease canopy fuels and, consequently, lower fire risk. The processionary pine
eats the needles before the spring, and it may lead to a completely defoliated pine
by the beginning of the growing season. This is already being observed in some
Mediterranean forests where, as a consequence of increasingly warm winters, moth
populations are increasing and expanding over larger areas. Areas that have suffered
intense pine processionary attacks are less likely to experience canopy fires because
those trees only host current year needles. Such attack does not kill the tree because
the buds are left intact, but the trees need to regrow their canopy with only current
year needles. Pines in uninfected areas can hold their needles over 3 or more years
so these drastic reductions in canopy area may act to diminish fire risk.
Fire might also act as feedback to catalyze the conversion into sparser vegetation.
As aridity increases, post-fire regeneration may be negatively affected by either
reduced rates of seedling emergence or by the resprout exhaustion syndrome as we
discussed in Chap. 8. Consequently, fire-induced deforestation could accelerate the
conversion of a PL- to a DL-fire type.

10.1.4 I ncreases in Forest Fire in Temperate and Boreal


Forests

One of the most likely consequences of climate change in the fire regime of temper-
ate and boreal forests will be an increase in fire activity (Millar and Stephenson
2015), albeit with different features across biomes and continents (Fig. 10.2).
10.1 Fire and Climate Change 185

Wildfires are currently rare in European temperate forests, where the fire-belt
was traditionally limited to the Mediterranean basin. Increasing aridity may trans-
form many of these areas into fire-prone ecosystems, a pattern that is already start-
ing to become apparent (Sommerfeld et al. 2018). Fire activity and burned area have
increased in temperate forests from North and South America, but this pattern is the
joint result of fire suppression, land management, and climate change (Westerling
et al. 2006). In the case of North America, forest fire activity has been associated
with El Niño Southern Oscillation and the Pacific Decadal Oscillation (Schoennagel
et al. 2005; Macias Fauria et al. 2011), and it also results from many decades of fire
suppression, which have resulted in large increases of available fuel. Wildfire in
temperate forests from South America, in contrast, is considered to have resulted
from a forest policy that encouraged large-scale plantations and subsequent increases
in fuel continuity (Gómez-González et al. 2018). While such forest plantations
report many different benefits, it is important to embed them within a fire-smart
landscape that prevents a large continuity of fuel accumulations.
There has been also an increase in fire activity in boreal forests, similarly result-
ing from increases in fuel availability after increased drought and increases in the
fire season (Flannigan et al. 2009). However, the pattern of wildfires varies mark-
edly between American and Eurasian boreal forests, despite having similar climates.
Such difference is considered to be driven by the contrasting physiological traits
from the tree species dominating these environments (Rogers et al. 2015). Crown
fires are common in North American boreal forests, which are forests dominated by
black spruce (Picea mariana), jack pine (Pinus banksiana), and white spruce (Picea
glauca), species with a low degree of self-pruning. In contrast, surface fires domi-
nate in Eurasian boreal forests, where Scots pine (Pinus sylvestris) and larch (Larix),
species with a high degree of self-pruning, dominate. A high degree of self-pruning
reduces the probability of a crown fire, because it reduces ladder fuels and breaks
the vertical continuity of the fuels. With climate warming, some models forecast an
increase in crown fire activity in Eurasian forests (Gauthier et al. 2015).
It is however unclear as to whether the increase in fire activity in temperate and
boreal ecosystems will be sustained over the twenty-first century or whether it will
be transient and a long-term decline in fire activity after fire-induced deforestation
may occur (Fig. 8.1). For instance, stand-replacing forests in Eurasian forests may
lead to a non-forest state as those species lack the capacity to regenerate after fire
(Shvidenko and Schepaschenko 2013; Lenton et al. 2008). Indeed, some modeling
studies indicate that continental steppe grasslands or open woodlands and grass-
lands could expand at the expense of boreal forests (Lenton et al. 2008). However,
other studies propose that forests could be maintained but that deciduous hardwoods
would substitute evergreen conifers (Tautenhahn et al. 2016).
It is important to remember that wildfires are just one of the different distur-
bances and stresses experienced by boreal forests. Increased frequency of droughts,
heat waves, and pathogens and pests could likely interact with increased wildfire
activity. In turn, dieback of boreal forest is particularly problematic from a global
climate perspective. Boreal ecosystems host large concentrations of carbon in peat
soils (Chap. 2) which could be released to the atmosphere under an increasing fire
186 10 Global Change, Pyrophysiology, and Wildfires

frequency. Some authors consider boreal forest decay as a tipping element in the
Earth’s climate system (Lenton et al. 2008), as vast amounts of greenhouse gases
could be emitted, with the potential to further reinforce climate change.

10.2  uture Wildfires, National Security, and Civil


F
Protection

10.2.1 Urbanization and Wildfires

Human ignitions globally play a small role in driving fire activity. That is, global
burned area would only increase by 3.5% if ignition limitations to fire activity were
to be removed (Kelley et al. 2019). However, they play a major role in some areas
of the world such as urban areas.
Urbanization increases the probability of ignitions, but it also declines forest
fuels which consequently protects areas from fire spread. Indeed, the relationship
between wildfires and population density is unimodal (Fig. 10.4). Wildfires are rare
in farm areas, where population density is low because management leads to low
fuel loads. As urbanization increases, there is an increase in ignitions. This leads to
a larger fire activity when urban areas sprawl into vegetated terrain or, conversely,
when forest areas sprawl into urban terrain. Further increases in urbanization and
very high population densities lead to declines in vegetation cover, which subse-
quently reduce fire activity (Syphard et al. 2007; Price and Bradstock 2014). Some
studies report that the effect of the degree of urbanization on burned area is globally
as important as the effect of population growth (Knorr et al. 2016).
The pattern of urbanization will be a crucial driver of future patterns of wildfire
(Radeloff et al. 2018). Wildfires within urban areas at the wildland-urban interface
(WUI) may be managed through fuel reduction treatments by creating a defensible

0.0020
Fire density (no. fires/km2)

0.0015

0.0010

0.0005

0
0 20 40 60 80 100
Population density (no. people/km2)

Fig. 10.4 The relationship between population and fire densities is unimodal. (Reproduced with
permission from Syphard et al. (2007))
10.2 Future Wildfires, National Security, and Civil Protection 187

space, that is, by reducing or completely eliminating woody vegetation over


30–50 m from the property (Syphard et al. 2014; Price et al. 2015). Additionally,
changes in construction materials and, in particular, in window preparation may
play a role as important as fuel reduction treatments (Syphard et al. 2017; Pastor
et al. 2019). It should be noted that there is great variation in construction materials
globally, so the Achilles heel may vary from country to country.
The areas that, to date, have suffered most losses from fourth- to fifth-generation
fires (Chap. 9) are concentrated in the most populous urban nuclei in Australia (e.g.,
Sydney) as well as in California. However, the Mediterranean Basin and other parts
of the world have also experienced such catastrophic fires. These catastrophes may
not be completely avoidable. Indeed, even houses with proper defensible space (i.e,
the separation between the property and the woody vegetation) are sometimes
burned, while the vegetation surrounding them survives. This is because the ember
loads may be more important than radiant or convective energy from the fire front
(Keeley and Syphard 2019). However, proper management may significantly reduce
the likelihood of catastrophes.
From the perspective of pyrophysiology, an aspect that deserves further attention
is the use of green firebreaks at the WUI (Curran et al. 2018). Green firebreaks could
be located outside the defensible space to separate the property from the surround-
ing vegetation. The rationale is that green firebreaks could help in reducing or stop-
ping fire spread and protect the property. Some authors propose that this tool has
additional benefits as they could be used to introduce species which might foster
local biodiversity. Property losses at the WUI result not only from radiation and
convection but also from firebrands. It would thus be beneficial if species within the
firebreaks show either low firebrand production or produce firebrands that show a
low spotting efficiency (Chap. 9). At any rate, the efficiency of green firebreaks
within the WUI, and which are the best species to be utilized, deserves further study
to assess its efficiency.

10.2.2  he Challenge of Mega-Fires, Increasing Fire Intensity


T
and Pyroconvection

Apart from changes in fire activity as such, perhaps the biggest challenge from a
human safety perspective is mega-fire activity. We have seen increases in fire inten-
sity over time such that six-generation fires, those with pyroconvection activity,
currently represent a major threat to life and property. In the fires occurred in 15
October 2017 in central Portugal, the blaze was able to burn 10,000 ha per hour for
13 h between 16:00 h and 05:00 h (Castellnou et al. 2018) (Fig. 9.1). This is an
unprecedented rate of spread for any European fire, and it was driven by pyrocon-
vective activity and the transition from pyrocumulus (PyroCu, smoke plumes
capped by cumulus clouds) to pyrocumulonimbus (PyroCb, a fire-triggered thun-
derstorm that creates erratic fire behavior). This is not the first time PyroCb occurs,
188 10 Global Change, Pyrophysiology, and Wildfires

and they may be common where fuel loads are large enough (Fromm et al. 2010).
However, pyroconvection often develops after 4–7 days of burning (Peterson et al.
2017), and not on the first day of fire as reported in Portugal.
There are two requisites for such plume-dominated fires: a high fire radiative
power and tropospheric conditions resembling those under dry thunderstorms
(Peterson et al. 2017). At present, pyroconvection activity seems to have been lim-
ited by fire radiative power, as the atmospheric conditions do not seem rare during
the fire season. Fire radiative power is a function of fuel loads, their structure, and
spatial pattern. Consequently, a major challenge for wildfire management during the
twenty-first century is to develop landscapes not prone to developing fire radiative
power large enough to drive PyroCb activity.
PyroCb is documented as a normal or, at least, as a phenomenon that is not rare
in, for instance, the wildfires from North America. Fromm et al. (2010) document
13 PyroCb events during June and July 2002, and Peterson et al. (2017) similarly
report 26 events during June and August 2013. A currently unresolved issue is
understanding what are the critical fuel loads that lead to a fire radiative power high
enough to produce either PyroCu or PyroCb under the right atmospheric conditions.
One of the surprises from the Pedrogao Grande fire in June 2017 was PyroCb activ-
ity in a surface fire. Understanding whether interspecific differences in heat capacity
and variations in factors such as moisture or ash content could alter the fire radiative
power at such scale is unlikely, but understanding critical fuel loads and the role of
dead biomass has not yet been examined.

10.3 Wildfires and Global Change

10.3.1 Wildfires and Plant Invasions

Plant invasions, whereby an exotic species is capable of becoming naturalized and


of expanding its distribution range after a human introduction, represent an increas-
ing thread to local biodiversity in many areas worldwide, and they exert profound
ecosystem modifications (Pyšek et al. 2012). One of the mechanisms underlying
plant invasions lies in their effects over the fire regime (Fig. 10.5) (Brooks et al.
2004). One of the best documented examples of this are invasive grasses, which
may increase the frequency and alter the intensity of fires, such that they take over
landscapes previously dominated by woody vegetation (Gaertner et al. 2014)
(Fig. 10.5a). The effect of invasive plants on fire regimes is most noticeable in Chile
and S Africa but also important in other fire-prone world areas (Moreira et al. 2019).
Grasses create thin biomass that dries annually and that subsequently increases
the load of highly available surface fuels. They are also able to create horizontal fuel
continuity to further ease fire spread. After the fire, many grasses are able to resprout,
and they recover quickly as they show little aboveground investments in aboveg-
round tissues. Such fast recovery implies an advantage over the native woody
10.3 Wildfires and Global Change 189

v­ egetation or over the native species lacking adaptations to survive under high fre-
quent fire. Over time, such disruption over the fire cycle favors the invasive grass at
the expense of native vegetation, which is not adapted to a high-frequency fire
regime (Fig. 10.5a), and also alters other ecosystem properties such as biogeochem-
ical cycles (Gaertner et al. 2014).
This process has been documented in many areas globally, including the invasion
of the C4 Schizachyrium condensatum and Melinis minutiflora in Hawai’i; of
Bromus tectorum in the Western USA; of African C4 grasses such as Hyparrhenia
rufa, Melinis minutiflora, or Panicum maximum in the Amazon; or of Cenchrus cili-
aris, Melinis minutiflora, or Pennisetum polystachion in Australia (D’Antonio and
Vitousek 1992).
Examples of woody plants that intensify the fire regime are more scarce, but they
have also been documented (Fig. 10.5b). For instance, invasion of Chromolaena
odorata in a South African savanna introduced a crown fire regime in an area where
surface fire was the native fire type (te Beest et al. 2011).
The introduction of fire-adapted species in fire-prone environments may also
alter ecosystem structure. Woody vegetation from the Everglades marshes, for
instance, had been naturally excluded by fire activity. The introduction of Melaleuca

(A) Fuel continuity


+ +
Grass establishment Fire frequency
+
Fuel availability
-
+

Native woody plant establishment

+ +
(B) Tree establishment Fuel ladder Fire intensity
-
+
Native vegetation establishment

(C) Fuel continuity


+ -
Tree establishment Fire frequency
-
Fuel availability
-
+
Native grass establishment

Fig. 10.5 Feedbacks between plant invasions and the fire regime, depending upon whether the
invasive is a grass (a), a tree that increases (b), or decreases (c) fire activity. Fuel availability indi-
cates dry foliage. Fuel continuity and fuel ladder refer to horizontal and vertical continuity,
respectively
190 10 Global Change, Pyrophysiology, and Wildfires

quinquenervia, a species with vigorous post-fire epicormic resprouting, transformed


these ecosystems into melaleuca forests (Schmitz et al. 1997).
The opposite process, where an invasive plant reduces fire activity, has also been
occasionally documented (Fig. 10.5c). Pine savannas from SE USA are maintained
by low-intensity surface fires. The invasive shrub Schinus terebinthifolius is fire
sensitive, but it managed to invade over the pine savannas by reducing fire fre-
quency. That is, areas with a high density of S. terebinthifolius showed low fire
frequency because the intense shading precluded understory fuel development
(Stevens and Beckage 2009). However, these results should be interpreted with cau-
tion because the results are based on a prescribed burn, and different results might
have been obtained under a higher-intensity natural fire.
Using a fire behavior model, van Wilgen and Richardson (1985) showed how the
very high leaf moisture content in Acacia saligna was able to reduce fire spread
when the species invaded a South African fynbos shrubland (Fig. 10.5c). Similarly,
Grace (1998) reported that the invasive tree Sapium sebiferum was able to transform
the fire-prone coastal prairies from Texas and Louisiana into a monoculture forest
which, consequently, showed much reduced fire frequencies (Fig. 10.5c).
Fire management may also exert significant impacts on plant invasions. Most
work in this area has focused in the western USA, an area with a significant ­presence
of grass and annual invasive plants. The most effective way to inhibit invasion has
been through fire suppression (Keeley 2006). Invasions have been promoted by
attempts to restore pre-European fire regimes. Pre-European plant composition has
changed, and such fire regimes now promote invasions. The use of prescribed burns,
the creation of firebreaks, and fuel reduction treatments through grazing have
favored invasions, particularly in shrubland ecosystems, but also in forests that were
susceptible to grass and annual invasions. In areas where annual or grass invasions
are less frequent, such as NW Spain, less significant impacts of fuel reduction treat-
ments occur over invasions (Fernández et al. 2019).

10.3.2 Wildfires and Biodiversity

Another result of global change is a major reorganization of species composition in


ecosystems globally. While the number of species has so far remained stable, sub-
stantial shifts in plant communities have occurred (Blowes et al. 2019), and models
forecast that up to 1 million species could undergo extinctions due to global change
(IPBES 2019). Wildfires have been present on land for 420 mya (Chap. 3), and they
have altered genetic diversity and shaped plant evolution (Chap. 4). Furthermore,
studies comparing diversification rates across plant phylogenies indicate that fire-­
prone lineages show higher diversification rates than non-fire-prone lineages (He
et al. 2019). It is thus to be expected that certain fire regimes will be necessary to
maintain and enhance biodiversity.
Plants are not adapted to fire per se but to a given fire regime. Thus species
exposed to a fire regime that matches their own requirements will be, in principle,
10.3 Wildfires and Global Change 191

favorable for the survival of that species. The more species with similar fire regime
requirements within an area, the more species that could benefit from a certain fire
regime. Any potentially positive effect of wildfire on biodiversity needs to be under-
stood from the perspective of plant adaptations to a fire regime. Any fire acting on a
tropical forest, for instance, will often lead to losses in biodiversity, because plants
in these ecosystems have evolved in the absence of fire activity (see Sect. 10.1.1).
One of the earliest hypotheses on how fire affects biodiversity is the intermediate
disturbance hypothesis (IDH), which states that species diversity is highest at inter-
mediate fire frequencies (Grime 1973). This hypothesis is a consequence of plants
being adapted to a certain fire regime. When fire return intervals are too short, or too
long, regeneration for some species may suffer, which consequently reduces plant
diversity.
Biodiversity is a complex process. On the one hand, it contains different compo-
nents, such as species richness, evenness (the relative proportion of the different
species), functional diversity (the diversity in functional attributes), or genetic
diversity, to name a few. On the other hand, biodiversity operates at local scales
(alpha diversity), across ecosystems (beta diversity), and over landscapes (gamma
diversity).
Landscape scale patterns of diversity may sprout from pyrodiversity. That is,
landscapes that consist of mosaics of different ages (time since last fire) often, but
not always, show higher diversity. A possible exception to this is old-growth ecosys-
tems, which might show enhanced biodiversity and species regeneration relies on
small-scale dynamics. The mechanisms explaining such a relationship include a
higher degree of environmental heterogeneity, the creation of new niches, a reduc-
tion of stress by reducing competition, as well as effects over trophic levels and a
temporal resource partitioning (He et al. 2019).
There are many studies on how fire affects biodiversity, as well as on the physi-
ological processes underlying such relationship (Kelly and Brotons 2017; He et al.
2019). However, studies on how biodiversity affects fire, and on the underlying
mechanisms, are less common.
Despite the many benefits of biodiversity, reductions in fire activity may not be
one of them. On the contrary, fire danger and crowning probability often increase
with structural complexity and continuity, which are results of higher species diver-
sity (Fig. 10.6). One of the mechanisms allowing for species coexistence is niche
complementarity, such that different species show different requirements and
resource uptake strategies. A result of such niche complementarity is that productiv-
ity and, consequently, fuel loads increase. One of the mechanisms explaining such
complementarity in forests is how canopy packing with interspecific competition
increased (Chap. 6). This effect increases the fuel ladder (vertical continuity) and
the canopy bulk density, a key parameter for crown fire rate of spread.
Niche complementarity may also originate from partitioning the hydrologic
niche. That is, different species may uptake water from different soil depths, which
allows for a more intensive water exploitation while diminishing plant competition.
However, niche redundancy might also occur, which would increase plant competi-
tion and decrease water availability. A recent review of the literature on how biodi-
192 10 Global Change, Pyrophysiology, and Wildfires

Fig. 10.6 Forests with


high structural, specific,
and functional diversity,
like the one in the picture,
show fuel ladders and are
more fire prone than
forests with simpler
structures (compared with
simple canopy structure in
Fig. 9.2). They might also
be more susceptible to
drought, which further
increases fire risk

versity affects drought responses in forests showed mixed responses, with some
studies reporting positive and others a negative or a neutral effect (Grossiord 2018).
The mechanisms underlying such responses are not yet fully elucidated. In addi-
tion to competition for water resources, facilitation provided by either shading or by
hydraulic lift (the passive movement of water from deep soil layers to shallower
depths) could also occur. Understanding physiological drought responses across
different tree species to understand biodiversity effects is thus an important question
to understand biodiversity-fire relationships. If biodiversity does exacerbate drought
effects, we could also observe earlier and more intense increases of the dead to live
ratio in biodiverse forests, as well as lower fuel moistures.
Although biodiversity may increase structural fire risk as well as fuel availability,
biodiversity may also enhance post-fire resilience. But it should be noted that this
effect is purely statistical. That is, the odds of having a species that may recover
after fire increase with the number of existing species. Despite this statistical effect,
many ecosystem functions may be recovered after a fire if one or more of the domi-
nant species are resprouters or show post-fire seeding.
The answer to this question likely lies on addressing wildfires as a process oper-
ating across stands, landscapes, and regions. The creation of fire-smart landscapes
requires a combination of areas where biodiversity conservation is a priority, with
others where we should prioritize economic development, climate change preven-
References 193

tion, tourism, and so on. There are many formidable challenges ahead in order to
better forecast future fire regimes and vegetation dynamics. Understanding the
plant-fire relations through a pyrophysiological approach may go a long way in that
direction.

References

Allen CD, Breshears DD, McDowell NG (2015) On underestimation of global vulnerability to tree
mortality and forest die-off from hotter drought in the Anthropocene. Ecosphere 6:129. https://
doi.org/10.1890/ES15-00203.1
Archer SR, Andersen EM, Predick KI, Schwinning S, Steidl RJ, Woods SR (2017) Woody
plant encroachment: causes and consequences. In: Briske DD (ed) Rangeland sys-
tems: processes, management and challenges. Springer, Cham, pp 25–84. https://doi.
org/10.1007/978-3-319-46709-2_2
Barlow J, Berenguer E, Carmenta R, França F (2019) Clarifying Amazonia’s burning crisis. Global
Change Biology in press (n/a). https://doi.org/10.1111/gcb.14872
Batllori E, Parisien M-A, Krawchuk MA, Moritz MA (2013) Climate change-induced shifts
in fire for Mediterranean ecosystems. Glob Ecol Biogeogr 22(10):1118–1129. https://doi.
org/10.1111/geb.12065
Blowes SA, Supp SR, Antão LH, Bates A, Bruelheide H, Chase JM, Moyes F, Magurran A,
McGill B, Myers-Smith IH, Winter M, Bjorkman AD, Bowler DE, Byrnes JEK, Gonzalez A,
Hines J, Isbell F, Jones HP, Navarro LM, Thompson PL, Vellend M, Waldock C, Dornelas M
(2019) The geography of biodiversity change in marine and terrestrial assemblages. Science
366(6463):339–345. https://doi.org/10.1126/science.aaw1620
Boer MM, Bowman DMJS, Murphy BP, Cary GJ, Cochrane MA, Fensham RJ, Krawchuk MA,
Price OF, Resco De Dios V, Williams RJ, Bradstock RA (2016) Future changes in climatic
water balance determine potential for transformational shifts in Australian fire regimes.
Environ Res Lett 11:065002
Boer MM, Resco De Dios V, Stefaniak EZ, Bradstock RA (2019) A hydroclimatic model for the dis-
tribution of fire on Earth. Biogeosci Discuss 2019:1–21. https://doi.org/10.5194/bg-2019-441
Bond WJ, Midgley GF (2000) A proposed CO2-controlled mechanism of woody plant invasion in
grasslands and savannas. Glob Chang Biol 6(8):865–869
Bond WJ, Midgley GF (2012) Carbon dioxide and the uneasy interactions of trees and savannah
grasses. Philos Trans R Soc Lond Ser B Biol Sci 367(1588):601–612. https://doi.org/10.1098/
rstb.2011.0182
Bradstock RA (2010) A biogeographic model of fire regimes in Australia: current and future impli-
cations. Glob Ecol Biogeogr 19:145–158. https://doi.org/10.1111/j.1466-8238.2009.00512.x
Brooks ML, D’Antonio CM, Richardson DM, Grace JB, Keeley JE, DiTomaso JM, Hobbs
RJ, Pellant M, Pyke D (2004) Effects of invasive alien plants on fire regimes. Bioscience
54(7):677–688
Cardil A, Molina DM, Oliveres J, Castellnou M (2016) Fire effects in Pinus uncinata Ram planta-
tions. For Syst 25(1):eSC06. https://doi.org/10.5424/fs/2016251-08919
Carnicer J, Coll M, Ninyerola M, Pons X, Sánchez G, Peñuelas J (2011) Widespread crown condi-
tion decline, food web disruption, and amplified tree mortality with increased climate change-­
type drought. Proc Natl Acad Sci 108:1474–1478. https://doi.org/10.1073/pnas.1010070108
Castellnou M, Guiomar N, Rego F, Fernandes PM (2018) Fire growth patterns in the 2017
mega fire episode of October 15, central Portugal. In: Viegas DX (ed) Advances in forest fire
research 2018. Imprensa da Universidade de Coimbra, Coimbra, pp 447–453. https://doi.
org/10.14195/978-989-26-16-506_48
194 10 Global Change, Pyrophysiology, and Wildfires

Charles-Dominique T, Midgley GF, Bond WJ, Scheiner S (2017) Fire frequency filters species
by bark traits in a savanna-forest mosaic. J Veg Sci 28(4):728–735. https://doi.org/10.1111/
jvs.12528
Cochrane MA, Schulze MD (1998) Forest fires in the Brazilian Amazon. Conserv Biol
12(5):948–950
Cochrane MA, Alencar A, Schulze MD, Souza CM, Nepstad DC, Lefebvre P, Davidson EA
(1999) Positive feedbacks in the fire dynamic of closed canopy tropical forests. Science
284(5421):1832–1835. https://doi.org/10.1126/science.284.5421.1832
Collins L, Boer MM, de Dios VR, Power SA, Bendall ER, Hasegawa S, Hueso RO, Nevado JP,
Bradstock RA (2018) Effects of competition and herbivory over woody seedling growth in a
temperate woodland trump the effects of elevated CO2. Oecologia 187(3):811–823. https://doi.
org/10.1007/s00442-018-4143-1
Cramer W, Guiot J, Fader M, Garrabou J, Gattuso J-P, Iglesias A, Lange MA, Lionello P, Llasat
MC, Paz S, Peñuelas J, Snoussi M, Toreti A, Tsimplis MN, Xoplaki E (2018) Climate change
and interconnected risks to sustainable development in the Mediterranean. Nat Clim Chang.
https://doi.org/10.1038/s41558-018-0299-2
Curran TJ, Perry GLW, Wyse SV, Alam MA (2018) Managing fire and biodiversity in the Wildland-­
Urban interface: a role for green firebreaks. Fire 1:3
Curt T, Mouillot F, Pellizzaro G (2013) Topic 13: measuring and modeling fuel change in rela-
tion to drought. In: Moreno JM (ed) Forest fires under climate, social and economic changes
in Europe, the Mediterranean and other fire-affected areas of the world. Lessons learned and
outlook, pp 30–31
D’Antonio CM, Vitousek PM (1992) Biological invasions by exotic grasses, the grass/fire
cycle, and global change. Annu Rev Ecol Syst 23(1):63–87. https://doi.org/10.1146/annurev.
es.23.110192.000431
Davis KT, Dobrowski SZ, Higuera PE, Holden ZA, Veblen TT, Rother MT, Parks SA, Sala A,
Maneta MP (2019) Wildfires and climate change push low-elevation forests across a critical
climate threshold for tree regeneration. Proc Natl Acad Sci U S A 116:6193–6198. https://doi.
org/10.1073/pnas.18151071165061/dryad.pc3f9d8
FAO (2011) The state of forests in the Amazon Basin. Congo Basin and Southeast Asia, Rome
Fernández C, Fernández-Alonso JM, Vega JA (2019) Effects of mastication of burned non-­
commercial Pinus pinaster Ait. trees on soil compaction and vegetation response. For Ecol
Manag 449:117457. https://doi.org/10.1016/j.foreco.2019.117457
Flannigan M, Stocks B, Turetsky M, Wotton M (2009) Impacts of climate change on fire activity
and fire management in the circumboreal forest. Glob Chang Biol 15(3):549–560. https://doi.
org/10.1111/j.1365-2486.2008.01660.x
Fromm M, Lindsey DT, Servranckx R, Yue G, Trickl T, Sica R, Doucet P, Godin-Beekmann S
(2010) The untold story of Pyrocumulonimbus. Bull Am Meteorol Soc 91(9):1193–1210.
https://doi.org/10.1175/2010bams3004.1
Gaertner M, Biggs R, Te Beest M, Hui C, Molofsky J, Richardson DM, Kühn I (2014) Invasive
plants as drivers of regime shifts: identifying high-priority invaders that alter feedback relation-
ships. Divers Distrib 20(7):733–744. https://doi.org/10.1111/ddi.12182
Gauthier S, Bernier P, Kuuluvainen T, Shvidenko AZ, Schepaschenko DG (2015) Boreal for-
est health and global change. Science 349(6250):819–822. https://doi.org/10.1126/science.
aaa9092
Giglio L, Randerson JT, van der Werf GR (2013) Analysis of daily, monthly, and annual burned
area using the fourth-generation global fire emissions database (GFED4). J Geophys Res
Biogeo 118:317–328. https://doi.org/10.1002/jgrg.20042
Gimeno TE, McVicar TR, O’Grady AP, Tissue DT, Ellsworth DS (2018) Elevated CO2 did not
affect the hydrological balance of a mature native Eucalyptus woodland. Glob Chang Biol
24(7):3010–3024. https://doi.org/10.1111/gcb.14139
References 195

Gómez-González S, Ojeda F, Fernandes PM (2018) Portugal and Chile: longing for sustainable
forestry while rising from the ashes. Environ Sci Pol 81:104–107. https://doi.org/10.1016/j.
envsci.2017.11.006
Grace J (1998) Can prescribed fire save the endangered coastal prairie ecosystem from Chinese
tallow invasion. Endangered Species Update, vol 15
Grime JP (1973) Competitive exclusion in herbaceous vegetation. Nature 242:344–347
Grossiord C (2018) Having the right neighbors: how tree species diversity modulates drought
impacts on forests. New Phytol. https://doi.org/10.1111/nph.15667
He T, Lamont BB, Pausas JG (2019) Fire as a key driver of Earth’s biodiversity. Biol Rev Camb
Philos Soc 94:1983–2010. https://doi.org/10.1111/brv.12544
Hódar JA, Castro J, Zamora R (2003) Pine processionary caterpillar Thaumetopoea pityocampa as
a new threat for relict Mediterranean Scots pine forests under climatic warming. Biol Conserv
110(1):123–129
IPBES (2019) Summary for policymakers of the global assessment report on biodiversity and
ecosystem services of the Intergovernmental Science-Policy Platform on Biodiversity and
Ecosystem Services. IPBES secretariat, Bonn, Germany
Karavani A, Boer MM, Baudena M, Colinas C, Díaz-Sierra R, Pemán J, de Luís M, Enríquez-­
de-­Salamanca Á, Resco de Dios V (2018) Fire-induced deforestation in drought-prone
Mediterranean forests: drivers and unknowns from leaves to communities. Ecol Monogr
88:141–169
Keeley JE (2006) Fire management impacts on invasive plants in the Western United States.
Conserv Biol 20(2):375–384. https://doi.org/10.1111/j.1523-1739.2006.00339.x
Keeley JE, Syphard AD (2019) Twenty-first century California, USA, wildfires: fuel-dominated
vs. wind-dominated fires. Fire Ecol 15(1). https://doi.org/10.1186/s42408-019-0041-0
Kelley DI, Bistinas I, Whitley R, Burton C, Marthews TR, Dong N (2019) How contemporary
bioclimatic and human controls change global fire regimes. Nat Clim Chang 9(9):690–696.
https://doi.org/10.1038/s41558-019-0540-7
Kelly LT, Brotons L (2017) Using fire to promote biodiversity. Science 355(6331):1264–1265.
https://doi.org/10.1126/science.aam7672
Knorr W, Arneth A, Jiang L (2016) Demographic controls of future global fire risk. Nat Clim
Chang 6(8):781–785
Krawchuk MA, Moritz MA (2011) Constraints on global fire activity vary across a resource gradi-
ent. Ecology 92(1):121–132. https://doi.org/10.1890/09-1843.1
Lenton TM, Held H, Kriegler E, Hall JW, Lucht W, Rahmstorf S, Schellnhuber HJ (2008) Tipping
elements in the Earth’s climate system. Proc Natl Acad Sci 105(6):1786–1793. https://doi.
org/10.1073/pnas.0705414105
Macias Fauria M, Michaletz ST, Johnson EA (2011) Predicting climate change effects on wild-
fires requires linking processes across scales. Wiley Interdiscip Rev Clim Chang 2(1):99–112.
https://doi.org/10.1002/wcc.92
McClaran MP, Ffoillot PF, Edminster CB (2003) Santa Rita experimental range:100 Years (1903 to
2003) of accomplishments and contributions. Conference proceedings, October 30–November
1, 2003. Tucson, AZ
Millar CI, Stephenson NL (2015) Temperate forest health in an era of emerging megadisturbance.
Science 349(6250):823–826. https://doi.org/10.1126/science.aaa9933
Moreira F, Ascoli D, Safford H, Adams M, Moreno JM, Pereira JC, Catry F, Armesto J, Bond
WJ, Gonzalez M, Curt T, Koutsias N, McCaw L, Price O, Pausas J, Rigolot E, Stephens S,
Tavsanoglu C, Vallejo R, Van Wilgen B, Xanthopoulos G, Fernandes P (2019) Wildfire man-
agement in Mediterranean-type regions: paradigm change needed. Environ Res Lett doi:https://
doi.org/10.1088/1748-9326/ab541e
Morgan JA, Milchunas DG, LeCain DR, West M, Mosier AR (2007) Carbon dioxide enrichment
alters plant community structure and accelerates shrub growth in the shortgrass steppe. Proc
Natl Acad Sci 104(37):14724–14729
196 10 Global Change, Pyrophysiology, and Wildfires

Moritz MA, Parisien M-A, Batllori E, Krawchuk MA, Van Dorn J, Ganz DJ, Hayhoe K (2012)
Climate change and disruptions to global fire activity. Ecosphere 3(6):art49. https://doi.
org/10.1890/es11-00345.1
Nobre CA, Sampaio G, Borma LS, Castilla-Rubio JC, Silva JS, Cardoso M (2016) Land-use and
climate change risks in the Amazon and the need of a novel sustainable development paradigm.
Proc Natl Acad Sci 113(39):10759–10768. https://doi.org/10.1073/pnas.1605516113
Pastor E, Munoz-Navarro JA, Caballero D, Àgueda A, Dalmau-Rovira F, Planas E (2019)
Wildland–Urban interface fires in Spain: summary of the policy framework and recommenda-
tions for improvement. Fire Technol. https://doi.org/10.1007/s10694-019-00883-z
Peñuelas J, Sardans J, Filella I, Estiarte M, Llusià J, Ogaya R, Carnicer J, Bartrons M, Rivas-­
Ubach A, Grau O, Peguero G, Margalef O, Pla-Rabés S, Stefanescu C, Asensio D, Preece
C, Liu L, Verger A, Rico L, Barbeta A, Achotegui-Castells A, Gargallo-Garriga A, Sperlich
D, Farré-Armengol G, Fernández-Martínez M, Liu D, Zhang C, Urbina I, Camino M, Vives
M, Nadal-Sala D, Sabaté S, Gracia C, Terradas J (2017) Assessment of the impacts of cli-
mate change on Mediterranean terrestrial ecosystems based on data from field experiments and
long-term monitored field gradients in Catalonia. Environ Exp Bot. https://doi.org/10.1016/j.
envexpbot.2017.05.012
Peterson DA, Hyer EJ, Campbell JR, Solbrig JE, Fromm MD (2017) A conceptual model for
development of intense pyrocumulonimbus in Western North America. Mon Weather Rev
145(6):2235–2255. https://doi.org/10.1175/mwr-d-16-0232.1
Price O, Bradstock R (2014) Countervailing effects of urbanization and vegetation extent on fire
frequency on the Wildland Urban Interface: disentangling fuel and ignition effects. Landsc
Urban Plan 130:81–88. https://doi.org/10.1016/j.landurbplan.2014.06.013
Price O, Borah R, Bradstock R, Penman T (2015) An empirical wildfire risk analysis: the probabil-
ity of a fire spreading to the urban interface in Sydney, Australia. Int J Wildland Fire 24(5):597.
https://doi.org/10.1071/wf14160
Pyšek P, Jarošík V, Hulme PE, Pergl J, Hejda M, Schaffner U, Vilà M (2012) A global assessment
of invasive plant impacts on resident species, communities and ecosystems: the interaction of
impact measures, invading species’ traits and environment. Glob Chang Biol 18(5):1725–1737.
https://doi.org/10.1111/j.1365-2486.2011.02636.x
Radeloff VC, Helmers DP, Kramer HA, Mockrin MH, Alexandre PM, Bar-Massada A, Butsic V,
Hawbaker TJ, Sn M, Syphard AD, Stewart SI (2018) Rapid growth of the US wildland-urban
interface raises wildfire risk. Proc Natl Acad Sci U S A 115:3314–3319
Resco de Dios V, Weltzin JF, Sun W, Huxman TE, Williams DG (2012) Windows of opportu-
nity for Prosopis velutina seedling establishment and encroachment in a semiarid grassland.
Perspect Plant Ecol Evol Syst 14(4):275–282
Resco de Dios V, Weltzin JF, Sun W, Huxman TE, Williams DG (2014) Transitions from grassland
to savanna under drought through passive facilitation by grasses. J Veg Sci 25:937–946
Rogers BM, Soja AJ, Goulden ML, Randerson JT (2015) Influence of tree species on continental
differences in boreal fires and climate feedbacks. Nat Geosci 8:228–234
Rosell JA (2016) Bark thickness across the angiosperms: more than just fire. New Phytol
211(1):90–102. https://doi.org/10.1111/nph.13889
Schmitz DC, Simberloff D, Hofstetter RH, Haller W, Sutton D (1997) The ecological impact of
nonindigenous plants. In: Simberloff D, Schmitz DC, Brown TC (eds) Strangers in paradise.
Island Press, Washington, DC, pp 39–61
Schoennagel T, Veblen TT, Romme WH, Sibold JS, Cook ER (2005) ENSO and PDO variabil-
ity affect drought-induced fire occurrence in rocky mountain subalpine forests. Ecol Appl
15(6):2000–2014. https://doi.org/10.1890/04-1579
Shvidenko AZ, Schepaschenko DG (2013) Climate change and wildfires in Russia. Contemp Probl
Ecol 6(7):683–692. https://doi.org/10.1134/S199542551307010X
Sommerfeld A, Senf C, Buma B, D’Amato AW, Despres T, Diaz-Hormazabal I, Fraver S, Frelich
LE, Gutierrez AG, Hart SJ, Harvey BJ, He HS, Hlasny T, Holz A, Kitzberger T, Kulakowski D,
Lindenmayer D, Mori AS, Muller J, Paritsis J, Perry GLW, Stephens SL, Svoboda M, Turner
References 197

MG, Veblen TT, Seidl R (2018) Patterns and drivers of recent disturbances across the temperate
forest biome. Nat Commun 9(1):4355. https://doi.org/10.1038/s41467-018-06788-9
Stephens SL, Collins BM, Fettig CJ, Finney MA, Hoffman CM, Knapp EE, North MP, Safford
H, Wayman RB (2018) Drought, tree mortality, and wildfire in forests adapted to frequent fire.
Bioscience 68(2):77–88. https://doi.org/10.1093/biosci/bix146
Stevens JT, Beckage B (2009) Fire feedbacks facilitate invasion of pine savannas by
Brazilian pepper (Schinus terebinthifolius). New Phytol 184(2):365–375. https://doi.
org/10.1111/j.1469-8137.2009.02965.x
Stevens N, Lehmann CE, Murphy BP, Durigan G (2016) Savanna woody encroachment is wide-
spread across three continents. Glob Chang Biol. https://doi.org/10.1111/gcb.13409
Syphard AD, Radeloff VC, Keeley JE, Hawbaker TJ, Clayton MK, Stewart SI, Hammer RB
(2007) Human influence on California fire regimes. Ecol Appl 17(5):1388–1402. https://doi.
org/10.1890/06-1128.1
Syphard AD, Brennan TJ, Keeley JE (2014) The role of defensible space for residential structure
protection during wildfires. Int J Wildland Fire 23(8):1165. https://doi.org/10.1071/wf13158
Syphard AD, Brennan TJ, Keeley JE (2017) The importance of building construction materials
relative to other factors affecting structure survival during wildfire. Int J Disaster Risk Reduct
21:140–147. https://doi.org/10.1016/j.ijdrr.2016.11.011
Tautenhahn S, Lichstein JW, Jung M, Kattge J, Bohlman SA, Heilmeier H, Prokushkin A, Kahl
A, Wirth C (2016) Dispersal limitation drives successional pathways in Central Siberian for-
ests under current and intensified fire regimes. Glob Chang Biol 22(6):2178–2197. https://doi.
org/10.1111/gcb.13181
te Beest M, Cromsigt JPGM, Ngobese J, Olff H (2011) Managing invasions at the cost of
native habitat? An experimental test of the impact of fire on the invasion of Chromolaena
odorata in a South African savanna. Biol Invasions 14(3):607–618. https://doi.org/10.1007/
s10530-011-0102-z
Turetsky MR, Benscoter B, Page S, Rein G, van der Werf GR, Watts A (2015) Global vulnerability
of peatlands to fire and carbon loss. Nat Geosci 8(1):11–14. https://doi.org/10.1038/ngeo2325.
http://www.nature.com/ngeo/journal/v8/n1/abs/ngeo2325.html. supplementary-information
Van Auken OW (2009) Causes and consequences of woody plant encroachment into western
North American grasslands. J Environ Manag 90(10):2931–2942. https://doi.org/10.1016/j.
jenvman.2009.04.023
van Wilgen BW, Richardson DM (1985) The effects of Alien Shrub invasions on vegetation struc-
ture and fire behaviour in South African Fynbos Shrublands: a simulation study. J Appl Ecol
22(3):955. https://doi.org/10.2307/2403243
Venter ZS, Cramer MD, Hawkins HJ (2018) Drivers of woody plant encroachment over Africa.
Nat Commun 9(1):2272. https://doi.org/10.1038/s41467-018-04616-8
Westerling AL, Hidalgo HG, Cayan DR, Swetnam TW (2006) Warming and earlier spring increase
western U.S. forest wildfire activity. Science 313(5789):940–943
Ying L, Han J, Du Y, Shen Z (2018) Forest fire characteristics in China: spatial patterns and
determinants with thresholds. For Ecol Manag 424:345–354. https://doi.org/10.1016/j.
foreco.2018.05.020
Zhou L, Tian Y, Myneni RB, Ciais P, Saatchi S, Liu YY, Piao S, Chen H, Vermote EF, Song C,
Hwang T (2014) Widespread decline of Congo rainforest greenness in the past decade. Nature
509(7498):86–90. https://doi.org/10.1038/nature13265
Index

A Aspect, 3, 5, 6, 22, 25, 36, 66, 68, 77, 119,


Abiotic drivers, 5 139, 144, 170, 187
Acacia saligna, 190 Atmosphere, 3, 8, 31, 32, 34, 36, 46, 78,
Active crown fires, 17, 99 167, 185
Adaptations, 6, 11, 15, 18, 22, 35, 54, 55, Australia, 2, 6, 9, 23, 37, 43, 46, 58, 79, 85,
57–62, 64, 65, 67–69, 106, 118, 156, 111, 128, 179, 187, 189
170, 172, 179, 180, 189, 191 Australian, 11, 26, 43, 57, 59, 110, 181
Adaptive, 15, 54 Autumn, 24
Aerial seed banks, 15, 44, 59, 139
Aerosol, 36, 37
Africa, 2, 23, 41, 42, 44, 46, 125, 182, 188 B
African, 26, 57, 189, 190 Banksia, 58, 108, 140
African savannas, 6, 42, 181, 189 Bark, 17, 20, 35, 60, 64–66, 68, 69, 105,
Albedo, 34, 36, 43 109–111, 118, 122–125, 127, 128, 137,
Algorithms, 22 145, 167, 179
Allocation, 55, 69, 96–99, 108, 127 Bark beetle, 108, 109
Alpine, 22 Bark density, 122
Amazon, 35, 45, 180, 189 Bark functions, 137
Amazonian Basin, 35 Bark shedding, 93
American, 179, 181, 185 Bark thickness, 8, 65, 66, 118, 122, 124, 171
Angiosperm, 56, 62, 105, 123, 129 Bark water content, 122
Anisohydric, 81, 144 Basal resprouting, 124, 125
Annuals = annual vegetation, 17 Bet hedging, 59–61
Anthropocene, 32, 45, 46 Bimodal fire season, 24
Anthropogenic fire regime, 44 Biodiversity, 9, 35, 187, 188, 190–193
Anti-flammability, 54, 55, 66–68 Biogenic volatile organic compounds
Apical dominance, 99, 100, 102 (BVOCs), 103, 104
Apical resprouting, 124 Biogeochemical cycling, 33
Apoplastic water, 81 Biogeophysical, 36
Archives, 24 Biomass, 3, 6–8, 18, 21, 22, 33, 36, 57, 62, 83,
Aridity, 6, 7, 59, 60, 64, 68, 110, 135, 93, 95, 97, 99, 100, 102, 109, 110, 119,
178–180, 184, 185 120, 156, 159, 162, 165, 170, 188
Arson, 5 Biomes, 15, 26, 27, 36, 41, 45, 63, 66, 101,
Ash, 37, 104–106, 127 125, 177, 179, 183–184
Ash content, 8, 57, 105, 166, 188 Biosphere, 3, 32, 38, 41
Biosphere-atmosphere interactions, 43

© Springer Nature Switzerland AG 2020 199


V. Resco de Dios, Plant-Fire Interactions, Managing Forest Ecosystems 36,
https://doi.org/10.1007/978-3-030-41192-3
200 Index

Branching pattern, 57 Chile, 9, 188


Broadleaves, 96 Chromolaena odorata, 189
Bromus tectorum, 172, 189 Civil protection, 9, 156
Browsers, 182 Climate, 7, 8, 31, 33, 34, 36, 40, 41, 43, 53,
Bud, 65, 94–96, 118, 120, 121, 124–125, 137 135, 142, 148, 156, 170, 171, 177–186
Bud bank, 64, 124, 125 Climate change, 9, 24, 34, 35, 43, 46, 126,
Bulb, 96, 125 134, 148, 170, 177, 179, 180,
Burn, 1, 5–8, 15, 17, 18, 22, 24–26, 35, 36, 44, 183–186, 192
45, 66, 68, 77, 78, 85, 88, 104, 105, Climate models, 34
111, 120, 127, 134, 156–159, 164–166, Climatic system, 31, 34
168, 169, 179, 181, 187, 190 Coastal prairies, 190
Burnt area, 156, 160 Collapse, 81, 158
Burnt patch size, 136 Combustion, 18, 35, 38, 46, 58, 76–78,
Byram, 18–21, 77 104–106, 118–123, 148
Combustion products, 54, 60
Communities, 31, 40, 66, 67, 133–135,
C 137–139, 142, 143, 145–148, 165, 169,
California, 24, 79, 85, 165, 187 171, 179, 190
Calluna vulgaris, 86 Community composition, 133–135, 168
Cambium, 65, 66, 118, 120–122, 137, 145 Compensatory growth, 37, 97
Cambium charring, 121, 122 Competition, 64, 66, 99–103, 110, 127, 128,
Cambium damage, 17, 127 139, 140, 142, 144, 145, 163, 164, 181,
Canada, 77 182, 191, 192
Canopy architecture, 8, 69, 95, 97, 99, 102, Combustibility, 66, 67, 167
118, 121, 163 Conduction, 5, 122, 123
Canopy base height (CBH), 95 Congo Basin, 45, 180
Canopy bulk density (CBD), 99–103, 137, Conifer, 26, 27, 40, 43, 44, 46, 56, 58, 59, 65,
163, 164, 169 68, 69, 82, 84, 99, 101, 102, 110, 128,
Canopy cover, 7 134, 164, 179, 180
Canopy depth, 95 Connectivity, 25, 96
Canopy moisture, 83 Consumability, 66, 67, 167
Canopy width, 95 Consumer, 43
Carbohydrates, 64, 97, 119, 122, 125, 126, Consumption, 118, 119, 123, 126, 144
142–144, 181 Convection, 5, 45, 120, 159, 187
Carbon allocation, 64 Cooling, 36, 37, 44, 159, 166
Carbon assimilation, 97, 98, 127, 143, 171 Corm, 125
Carbon budget, 8 Cretaceous, 41, 58, 61, 62
Carbon emissions, 45 Cretaceous-paleocene extinction event, 57
Carbon gain, 95, 97, 127, 170 Critical flammability, 5, 87, 88
Carboniferous, 39, 40, 58 Crown = canopy, 17, 99–102
Carbon sink = C sink, 34 Crown fire = crowning, 15–18, 21, 26, 40, 44,
Carbon source = C source, 145 46, 60, 67, 69, 84, 93–95, 99, 101, 105,
Carbon starvation, 119, 120, 142–144 107–110, 119, 120, 123, 137, 138, 147,
Carbon transport, 119 157, 158, 163, 185, 189, 191
Catastrophic wildfires, 7, 9, 40, 43, 45, 79 Crown fuel, 16, 17, 108
Cavitation, 99, 107, 108, 120, 122–123, 127, Crown scorch, 20, 120, 127, 164
137, 143, 144 Crown volume, 101, 121
Cell wall deformation, 123 C stocks, 34
Cell wall elasticity, 81 Cupressaceae, 58
Cenchrus ciliaris, 189 Cured, 19
C4 grasses, 189 Curing, 95
Chamaephytes, 96 Cuticular conductance, 81, 144
Index 201

D Ecosystems, 1, 7, 8, 10, 11, 15, 17, 24, 26,


Dead biomass, 83, 138, 148, 164, 184, 188 34–37, 40, 43, 58, 60, 64, 67, 75, 85,
Dead fine fuel moisture model, 78, 79 88, 93, 94, 102, 104, 110, 111, 126,
Dead floret retention, 108 134, 147, 148, 155, 156, 161, 165, 171,
Dead fuel, 5, 76, 78, 79, 83, 85, 86, 93, 177–179, 181, 183, 185, 188–192
94, 106–108 Ecosystem structure, 31, 32, 103, 189
Dead fuel moisture content, 79, 86 Elephants, 182
Dead leaf retention, 108 Embolism, 64, 68, 69, 99, 137
Dead matter, 57, 83, 84, 94, 166 Emissions, 8, 31, 34–36, 40, 45, 46, 103, 104,
Decomposition, 33, 93–95, 105, 110–111, 166 162, 165
Decorticating bark, 96, 109, 110 Emmenanthe penduliflora, 60
Defense, 57, 68, 97, 105, 128, 164 Energy release, 17–19, 86, 104, 158
Defoliation, 16, 108, 118, 120–122, 126, 127, Environmental pressure, 54, 55, 57
183, 184 Environmental stress, 22, 82
Deforestation, 27, 35, 36, 134, 135, 138, 179, Epicormic, 96, 124, 125
180, 184, 185 Epicormic bud, 57, 62
Degradation, 34, 45, 134, 147 Epicormic resprouting, 124, 190
Dendrochronology, 23, 168 Erosion, 8, 33, 35, 37, 38, 139, 147
Dependent crown fires, 17 Eucalyptus, 103, 109, 124
Deposition, 93–95, 110–111, 167, 169 Eucalyptus regnans, 37
Desert, 6 European, 11, 24, 25, 43, 44, 60, 100, 181,
Desertification, 183 185, 187
Development, 3, 10, 15, 18, 19, 41, 42, 61, Event-dependent hypothesis, 142
93–103, 105, 107–111, 122, 141, 144, Everglades, 189
156, 161, 162, 167, 169, 170, 192 Evolution, 8, 10, 11, 31, 41, 53–57, 59–62,
Devonian, 1, 39 66–69, 137, 156, 171, 182–184
Diffuse-porous species, 127 Exaptations, 54, 55, 67
Dinosaurs, 41 Extinction capacity, 9, 17, 161
Direct attack = direct fire attack, 20, 21, 157 Extinctions, 25, 40, 41, 43, 145, 148, 190
Direct fire effects, 118, 120 Extreme fire behavior, 17, 111
Disturbance, 54, 57, 62, 69, 93, 95, 97, 135, Extremely large wildfire, 8, 169
165, 179, 185, 191
Diversity, 22, 53, 94, 95, 101, 111, 124, 127,
165, 190–192 F
Drought deciduous, 107, 108, 144, 183 Facultative resprouter, 146
Drought-induced leaf senescence, 84, 107, Farming, 9, 42–45
108, 183 Fatalities, 16, 46, 79, 156
Drought = water stress = water scarcity, Fear, 10, 161
25, 61, 87 Fear trap, 10
Dry, 5, 6, 9, 24, 25, 40, 42, 44, 59, 69, 77, Feedback, 3, 32–34, 36, 38, 41, 44, 127, 128,
80–82, 84–86, 88, 105, 107, 108, 110, 133–148, 184, 189
123, 141, 142, 144, 164, 172, 188, 189 Fiber saturation point, 77
Dryness-limited fire domain, 75 Fine dead fuel, 5, 76
Dry periods, 16 Fire activity, 3, 4, 6, 7, 15, 24, 31, 33, 34,
Duff, 17, 18, 105 38–44, 46, 54, 56, 57, 59, 60, 62, 63,
66, 68, 69, 76, 78, 79, 85–88, 155, 166,
168, 172, 177–181, 183–187, 189–191
E Fire-adapted, 57, 189
Early-to-rise hypothesis, 141 Fire avoider, 69
Earth’s history, 1, 53 Fire behavior, 1, 3, 11, 18, 68, 77–79, 84, 93,
Earth system, 9–11, 31–47, 134 100, 103, 104, 106, 107, 109, 111, 137,
Ecological perspective, 3 138, 156–158, 166, 167, 169, 187, 190
Ecophysiology, 1–3, 159, 161, 162, 169–172 Firebrand, 96, 109, 159, 187
202 Index

Firebreaks, 9, 156, 157, 159, 161, 166, 182, Fire spread, 5, 16, 17, 20, 33, 76, 77, 85–87,
187, 190 99, 100, 104, 105, 107, 109–110, 158,
Fire consumption, 16 159, 161, 162, 166, 168, 186–188, 190
Fire cycle = fire rotation period, 22, 23 Fire statistics, 25
Fire danger, 10, 24, 75–88, 109, 169, 191 Fire-stick, 42, 43
Fire danger indices, 75–77 Firestorm, 20, 21, 41, 109
Fire deficit, 6 Fire suppression, 6, 9, 17, 45, 155, 156,
Fire domains, 75 185, 190
Fire embracer, 69, 146 Fire suppression policy, 9, 25, 27, 35
Fire environment, 5, 61, 168 Fire syndromes, 26
Fire frequency, 6, 15, 22–24, 44, 179, 181, Fire tolerant, 68, 69
182, 185–186, 190, 191 Fire types, 15–18, 23, 60, 67, 79, 86, 95, 156,
Firefighting, 17, 46, 155–157 158, 159, 167, 169, 178, 179, 189
Fire-forage, 42 Fire use, 42–44
Fire front=front, 19, 77, 109, 122, 187 Fire weather, 5, 6
Fire generations, 156, 157, 166 Fire weather index (FWI), 77, 79
Fire indices, 79 Fire = wildfire = wildland fire, 23, 75–88, 103
Fire intensity = intensity Fitness, 54, 55, 67, 107
Fire intervals, 23, 63, 144 Flame, 20, 31, 35, 157, 166, 180
Fire inventories, 24 Flame depth = depth of the flame, 5
Fireline, 19, 20, 107 Flame height = height of the flame, 16,
Fireline intensity = Byram’s intensity = 122, 168
Byram’s fireline intensity, 19 Flame length = length of the flame, 5
Fire management, 1, 11, 43, 46, 190 Flame transport, 5
Fire mosaic, 25 Flaming, 76, 77
Fire occurrences, 22, 24, 77 Flammability, 5, 6, 8, 11, 54, 55, 57, 66–68,
Fire paradox = fire-fighting trap = paradox of 93, 103–108, 111, 137, 138, 156,
extinction, 9 158, 166–168
Fire prevention, 10, 156, 157, 159–162, Flammability thresholds, 5, 87, 88
164, 169 Flowering, 57, 61, 128, 129, 146
Fire-prone environments, 11, 54, 56, 58, Fluid mechanics, 3
61, 66, 189 Follicle, 58
Fireproof buildings, 158 Forest, 3, 7, 16, 17, 19, 22–24, 26, 27, 35, 37,
Fire propagation, 5, 17, 33, 161, 163 39, 40, 45, 75, 77, 79, 83–85, 95, 96,
Fire protection, 65, 137 110, 121, 123, 126–128, 133–148, 156,
Fire radiative energy (FRE), 122 159, 162, 177–180, 183–186, 190–192
Fire radiative power (FRP), 19 Forest fuel, 186
Fire records, 38 Forest planning, 20, 161
Fire refugia, 140 Forestry, 1
Fire regime, 3, 6, 8–11, 15–21, 23–27, 38, Forest structure, 35
40–46, 54, 56, 60, 63, 66, 68, 95–96, Fossil fuel emissions, 34, 35
118, 128, 133, 134, 139, 142, 144, 159, Fossil record, 38, 58
160, 165, 168, 177, 180, 181, 183, 184, Four-switch concept, 5, 6
188–191, 193 Fraction of absorbed photosynthetically active
Fire return interval, 22–24, 27, 66, 67, 95, 141, radiation (faPAR), 22
145, 191 Fuel, 1–7, 9, 10, 16–20, 24, 25, 34, 36, 38, 39,
Fire science, 15, 96 41–43, 46, 68, 75–88, 93–111, 125,
Fire season, 24, 27, 79, 83, 84, 86–88, 95, 107, 134, 148, 156–161, 164–169, 177–180,
108, 165, 170, 185, 188 182–185, 192
Fire seasonality, 15, 24 Fuel availability, 5, 6, 45, 126, 137, 179, 180,
Fire severity = severity, 21, 22, 37, 107, 119 183, 185, 189, 192
Fire size, 25–27 Fuelbed, 16, 107
Fire smart landscape, 185 Fuel build-up, 17, 24, 156, 157, 160, 177
Index 203

Fuel continuity, 9, 156, 158, 185, 189 Height, 16, 20, 57, 95, 96, 102, 109, 122, 137,
Fuel development, 3, 10, 190 168, 181, 182
Fuel ladder, 16, 17, 162, 164, 189, 191 Helenium aromaticum, 56
Fuel layers, 15, 16, 97 Hemicryptophytes, 96
Fuel load, 3, 5, 6, 9, 10, 16, 17, 20, 69, 93, 95, Herbaceous vegetation = herbs, 16, 145
97, 100, 101, 103, 105, 107, 108, 111, Herbivores, 7, 8, 43, 59, 65, 66, 104, 138, 141,
157, 159–164, 169, 177, 179, 184, 186, 144, 181, 182
188, 191 Hominins, 41, 42
Fuel management, 1, 169 Horizontal fuel continuity, 17, 157, 188
Fuel moisture, 2, 10, 33, 75–79, 82–86, 88, 93, Hormones, 99
95, 106–108, 137, 162, 163, 183, 192 Human activity, 6, 41, 43, 45, 46, 179, 180
Fuel reduction, 165 Human evolution, 31, 42
Fuel reduction treatments, 9, 21, 157, 160, Human pressure, 6
161, 186, 187, 190 Humans, 1, 5–7, 23, 27, 31, 32, 38, 40–43, 45,
Functional types, 61, 64, 106, 107, 118, 124 46, 180, 186–188
Fynbos, 190 Humidity, 5, 76, 77, 121, 158, 163, 166
Hydraulic architecture, 171
Hydraulic conductance, 69, 119
G Hydraulic efficiency, 99
Geological scales, 32, 38 Hydraulic failure, 84, 143
Geophytes, 96, 125, 129 Hydrophobic layer, 37, 38
Geosphere, 31 Hydrosphere, 31
Germination, 56, 60–62, 141, 144, 165 Hyparrhenia rufa, 189
Germination cues, 8
Global, 1, 2, 6, 8, 11, 16, 23, 26, 35, 40, 41,
177, 180, 181, 185, 186, 188–193 I
Global change, 11, 42, 177–193 Ignitability, 66, 67, 167
Global synthesis, 43 Ignite, 17, 42, 76, 77, 86, 104, 111
Global warming, 162, 183 Ignition, 5, 6, 18, 38, 42, 66, 77, 85, 95, 103,
Grasslands, 7, 16, 19, 26, 27, 35, 84, 134, 185 104, 160, 180, 186
Grass-tree interactions, 182 Ilex aquifolium, 99
Grassy fuels, 19, 181, 182 Imbibition, 60
Grazers, 45, 182 Impact = fire impact = ecological
Greenhouse gas, 45, 165 impact, 18
Greenhouse gas balance, 31 Independent crown fires, 17, 157
Ground fire, 16, 18, 21, 35, 46, 119, 123 Indirect fire effects, 118, 120
Growth, 2, 3, 5, 24, 35, 42, 64, 97–100, 102, Indonesia, 35, 180
103, 118, 120, 125, 127–128, 162–165, Inertinite, 38
167, 169, 170, 181, 186 Infra-red, 19
Growth form, 94 Inner bark, 57, 66, 128
Growth rates, 95–99, 144, 160, 163, 166, 167, Insect attacks, 128, 164, 183
170, 181 Intermediate fire-productivity hypothesis, 6
Gymnosperm, 123 Intra-specific variation, 103, 121
Invasives = plant invasions, 171, 188–190
Inventories, 21, 22
H Isohydric, 81, 144
Habitat, 38, 43, 148 Isoprene, 103
Haemodoraceae, 61
Hard coated seed, 62
Hawai’i, 189 J
Health, 46, 126, 155, 156 Jurassic, 40
Heat, 16, 19, 31, 56–58, 60, 77, 79, 86, 122,
123, 139, 148, 168, 169, 185, 188
Heat content, 95 K
Heat shock, 54, 61 Karrikins, 61
204 Index

L Management goals, 155


Lagged responses, 118 Mastication, 165, 170
Land abandonment, 9, 17, 157, 162, 183 Masting, 59, 129
Land management, 6, 7, 9, 11, 15, 31, 44, 156, Maternal effects, 62
157, 161, 185 Mechanisms, 1, 3, 10, 11, 32, 54–58, 61, 67,
Landscape, 3, 5–7, 9, 11, 32, 36, 38, 41–47, 69, 81, 88, 99, 108, 118, 119, 123, 126,
66, 78, 82, 103, 104, 111, 138, 141, 129, 134, 137, 141, 145, 161, 167, 168,
155–163, 166–169, 182, 188, 191, 192 171, 181, 182, 188, 191, 192
Land use, 5, 6, 18, 44, 45, 179, 180 Mediterranean, 6, 11, 18, 24, 26, 27, 59,
Large wildfire, 4, 5, 9, 24, 25, 40, 148, 61–64, 79, 82, 84–88, 124–127, 134,
156, 162 135, 138–141, 143, 145, 146, 148, 156,
Latent heat, 78, 85, 166 157, 160–162, 178, 183–184, 187
Leaf age, 83 Mediterranean Basin, 9, 11, 24, 46, 108, 162,
Leaf angle, 57 183, 185
Leaf area, 37, 57, 97, 98, 102, 108, 137 Mediterranean climates, 62, 63
Leaf area density, 101, 102 Mega-fire, 187
Leaf area index, 101 Meiosis, 57, 58
Leaf clumping, 100, 101 Melaleuca quinquenervia, 189–190
Leaf economics, 57, 68 Melinis minutiflora, 189
Leaf length, 101, 107 Meristem, 64–66, 99, 124
Leaf mass area, 95 Meristem protection, 57
Leaf moisture, 81, 95, 105, 107, 166, 190 Metal concentrations, 37
Leaf-scale flammability, 66 Metal oxides, 58
Leaf senescence, 84, 105, 107–109, Meteorology, 3, 159
137, 169–170 Middle wave Infra-Red (MIR), 22
Leaf shape, 95 Mineral content, 95, 104–106, 166
Leaf shedding, 84, 93, 108 Mineralization, 8, 139, 147
Leaf size, 57, 106, 107, 111, 118 Miocene, 41, 59, 62, 63
Legacy effects, 118, 120, 143 Mires, 18
Li-Dar, 101, 103 Mixed forest, 101, 135, 147, 180
Light-demanding, 99 Moisture, 5, 8, 33, 57, 68, 69, 75–88, 95, 106,
Lightning, 38, 41, 42 108, 141, 166, 169, 177, 188
Light tolerance, 57, 68 Montane, 26, 27, 40, 69, 128, 146
Lignotuber, 57, 125, 181 Mortality, 21, 31, 41, 46, 110, 117, 118,
Linear firebreak, 157 120–123, 127, 128, 133–135, 137, 144,
Linear infrastructures, 157, 159, 166 167, 169, 179, 182, 183
Liquidambar styraciflua, 99 Mosaic, 9, 25, 96, 191
Litter, 5, 16, 20, 21, 104, 105, 107, 110, 141, Mountainous regions, 17, 127, 179
163, 164, 166, 182 Mutagenic, 55–58
Litterfall, 110 Mutation, 58
Litter fuels, 16, 107 Mycorrhiza, 127
Live fuel, 5, 76, 85, 88, 94, 103, 104, 107, Myrtaceae, 57
137, 162, 166, 169
Live fuel moisture content, 76, 79, 80, 83, 106
Live/dead ratio, 95 N
Logging, 35, 110, 170 Near Infra-Red (NIR), 22
Louisiana, 190 Neolithic, 43
Niche, 137, 138, 148, 162, 171, 191
Nitrate, 37
M Nitrogen, 57, 97, 98, 170
Machine learning, 22 Nonstructural carbohydrates, 81, 106, 125
Maladaptive, 15, 54, 62 North America, 9, 108, 185, 188
Mammals, 43 Nutrient availability, 8, 127, 139, 147, 181
Index 205

Nutrient losses, 8, 38, 147 Photosynthates, 65, 69, 103, 137, 142
Nutrient recycling, 105 Photosynthesis = assimilation = C assimilation
= CO2 assimilation, 34, 35, 97–100,
125, 127, 143, 171
O Photosynthetic tissues, 21
Oaks, 43, 103, 126, 135, 144, 146–148, 184 Physiognomic changes, 7, 34, 43
Obligate resprouters, 8 Physiognomy, 36, 43
Obligate seeders, 63, 146 Physiological perspective, 3
Oils, 103, 180 Picea brachytyla, 126
Open forests, 17 Pines, 20, 21, 44, 78, 84, 99, 100, 105, 108,
Organic burial, 33 109, 127, 128, 134, 135, 137, 139, 140,
Organic carbon = organic C, 32, 33 145, 146, 148, 162–164, 170, 184,
Organic matter, 18, 32, 34, 36, 37, 39, 185, 190
61, 76, 180 Pinus, 44, 58, 65, 68, 103, 104, 122, 145,
Organic nitrogen, 37 162, 185
Osmotic potential, 81 Pinus aristata, 69
Outer bark, 57, 66, 67, 122 Pinus attenuata, 69
Oxygen = O2, 8, 18, 31–33, 38, 39, 41, 76, 77, Pinus halepensis, 84, 107, 108, 121, 135, 139,
156, 170 140, 162
Pinus nigra, 128
Pinus pinaster, 121, 148
P Pinus ponderosa, 129
Packing ratio, 95, 107, 166 Pinus sylvestris, 185
Paleocene-Eocene boundary, 63 Pinus uncinata, 69, 100
Paleogene, 65 Plant canopies, 5, 83, 99
Paleozoic, 38–40 Plant cover, 33, 37, 142
Pangea, 40 Plant death, 93, 107–109, 118, 123, 133
Panicum maximum, 189 Plant ecology, 1
Particulate matter, 31, 46 Plant evolution, 53, 54, 66, 69, 190
Passive crown fires, 17 Plant flammability, 66
Pastoralism, 42, 45 Plant performance, 117–129
Pathogens, 65, 104, 105, 107, 108, 126, 141, Pliocene, 41
184, 185 Plume = heat column, 122, 123
P cycle, 31, 33 Pollen, 23
Peat, 40, 111, 180, 185 Polycyclic aromatic hydrocarbons, 58
Peat fires, 36, 46, 76, 180 Polyploidy, 57
Pennisetum polystachion, 189 Populations, 6, 9, 42, 44, 45, 59, 69, 118, 121,
Pericarp, 56 126, 133–146, 184, 186
Permian, 39, 40 Portugal, 16, 79, 156, 158, 187, 188
Persistence, 6, 57, 62–64, 96, 126, 171 Post-fire establishment, 64, 141
Perturbation, 7, 33, 38, 43, 60, 68, 97, Post-fire success, 54, 62
126, 171 Post-fire vegetation dynamics, 10, 136
Pests, 107, 108, 126, 138, 141, 162, 165, Precipitation, 35, 37, 45, 59, 60, 64, 65, 77,
184, 185 78, 110, 138, 142
Phanerophytes, 96 Predawn water potential, 82, 144
Phanerozoic, 39 Pre-programmed leaf senescence, 84, 107
Phenocam, 83 Productive, 63, 64, 68
Phenology, 57, 68, 95, 137 Productivity, 6, 7, 63, 64, 69, 110, 138,
Phloem, 69, 118–122, 142 160–163, 177, 178, 192
Phloem flow, 143 Productivity-limited type fire, 178, 179, 183
Phosphate, 33 Propagule, 133, 136–145, 179
Phosphorous, 37, 57, 98 Proteaceae, 108
Photoperiod, 57 Protopitys buchiana, 65
206 Index

Proxy, 18, 38 Respiration, 109, 110, 117, 125, 143


Pruning, 17, 170 Resprouter, 43, 61, 63–65, 67, 81, 82, 97, 98,
Pterospartum tridentatum, 86 121, 124–126, 135, 138, 144, 146,
Putting out fires, 9 148, 192
Pyriscence, 59 Resprouting, 2, 6, 18, 26, 37, 54, 57, 62–64,
Pyroconvection, 157, 169, 179, 187, 188 81, 96, 97, 118, 123–127, 129, 135,
Pyrocumulonimbus = pyroCb, 158, 187, 188 137–139, 144, 165, 171, 181
Pyrocumulus = pyroCu, 158, 187 Resprouting lineages, 62, 63
Pyrodiversity, 156, 191 Restionaceae, 61
Pyrogardening, 158 Retardant, 105
Pyrogenic carbon = soot, 36 Rhizome, 96, 125
Pyrogeography, 6 Ring-porous species, 127
Pyrohabitat, 43 Root, 37, 58, 62, 64, 81, 97–99, 107, 108, 118,
Pyrolysis, 76 119, 122, 123, 125, 137, 141, 142, 144,
Pyrome, 26, 27, 45 162, 181
Pyrophysiology, 1, 3–6, 10, 11, 88, Root area, 120
155–172, 177–193 Rooting pattern, 81, 82
Rooting system, 64, 81
Rosmarinus officinalis, 140, 148
Q Ruderals, 145, 165
Quaternary, 41–45 Run-off, 34
Quecus cerrioides, 147 Rural exodus, 9, 35, 156
Quercus faginea, 147
Quercus ilex, 80, 126
S
Safe sites, 141, 142
R Salvage logging, 170
Radiation, 5, 20, 21, 36, 98, 120, 122, 139, Sap flow, 127
163, 187 Sapium sebiferum, 190
Radiative forcing, 37 Sap transport = water transport, 119, 120,
Rainfall, 7, 82, 141, 144, 181 122, 123
Rainforests, 6, 26, 27, 40, 110 Savanna grasslands, 62, 63
Rapid evolution, 56 Savannas, 7, 17, 22, 35, 41, 45, 63, 125, 148,
Rate of spread = spread rate, 33, 94, 99 180–182, 190
Raunkiaer, C., 94 Savannization, 35, 180
Reaction intensity = Rothermel’s intensity = Schinus terebinthifolius, 190
Rothermel’s reaction intensity, 19, 105 Schizachyrium condensatum, 189
Recruitment, 6, 54, 57–61, 66, 69, 133, 136, Scorch, 118, 120, 137
139, 142, 146, 171, 182 Seasonality, 15, 24
Reductionism = reductionist discipline, 3 Seed, 56–64, 67
Reforestation, 44 Seed banks, 6, 18, 26, 66, 96, 136–140, 165
Regeneration, 2, 6, 8, 22, 24, 26, 37, 55, 60, Seed dormancy, 60
63, 64, 67, 69, 80–82, 93, 96–99, 121, Seed germination, 57, 60–62, 134, 136
122, 128, 134, 135, 137, 139, 140, 144, Seeder, 8, 61, 63–65, 67, 82, 120, 135, 137,
146, 156, 171, 184, 191 138, 140, 144, 165
Relative humidity, 5, 79 Seeding, 62–64, 123, 192
Relief, 5 Seeds, 8, 15, 81, 96, 97, 129, 134, 136–141,
Remote sensing, 19, 21, 22, 24, 82, 83 144–147, 162
Reproduction, 53–54, 59, 69, 97, 118, Selection, 55, 66, 67
128–129, 171 Self pruning, 26, 69, 95, 109, 166, 185
Residence time, 19, 107, 122, 168, 180 Sensible heat, 34, 36
Resilience, 43, 126, 133–145, 192 Serotiny, 8, 26, 57–60, 65, 139, 171
Resins, 57, 58, 103, 104, 164 Settlements, 9, 43, 44, 181
Resistance, 33, 68, 69, 99, 123, 134, 136, 137, Shade tolerance, 57, 68, 98
146, 148, 171 Shoot age, 83
Index 207

Shoot-to-root, 98, 99 Subtropical savannas, 40, 181


Shrublands, 3, 17, 19–21, 24, 26, 27, 79, 82, Succession, 133–148
86, 134, 143, 144, 146–148, 178, 179, Summer, 24, 60–62, 81, 86, 87, 107, 123, 141,
184, 190 142, 172
Shrubs = shrubby vegetation, 16 Summer fires, 107
Silica, 105 Supernova, 41
Silurian, 1, 39 Surface area to volume ratio, 95
Sink activity, 97 Surface fires, 16, 17, 21, 23, 26, 40, 60, 61, 64,
Slash, 19, 44 65, 67, 68, 76, 95, 105, 107, 119, 120,
Slash and burn, 42, 44, 45 122, 128, 134, 137, 145, 146, 157, 166,
Slope, 5, 20, 45, 147, 159, 169 185, 188–190
Smoke, 31, 35, 45–47, 54, 58, 60–62, 77, Surface fuel, 17, 162, 164, 184, 188
165, 187 Surface temperature, 36, 123
Smoldering, 18, 46, 76, 77, 105, 120, 123 Survival, 2, 10, 53–55, 57, 65, 69, 96, 118,
Snow, 36 120, 121, 127–128, 135–137, 141–145,
Social, 22, 156 162, 171, 191
Society, 9 Sustainability, 66, 167
Socio-ecological, 22 Sydney, 79, 187
Socioeconomic problems, 11 Symbiotic relationships, 81, 141, 145
Soil, 8, 18, 34, 35, 37, 38, 45, 58–61, 64, 76, Symplastic water, 81
80, 81, 99, 110, 123, 125, 135–140, Synoptic conditions, 88
142, 147, 162, 167, 171, 181, 182, Synoptic weather patterns, 5
185, 192
Soil depth, 65, 191
Soil fertility, 8, 147 T
Soil organic matter, 8, 18, 37, 147 Temperate, 11, 26, 27, 36, 79, 84, 85, 125,
Soil properties, 8 179, 184–186
Soil seed banks, 60, 134 Temperature, 5, 16, 20, 24, 36, 40, 54, 61, 65,
Soil type, 7 76, 77, 79, 104, 106, 110, 118,
Source activity, 97 121–123, 125, 141, 142, 146, 158, 170
South Africa, 9, 42, 58 Temperature stress, 54
Spain, 24, 25, 43, 44, 85, 87, 121, 128, 156, Terpenes, 103
184, 190 Terrorism, 156
Spark, 5, 122, 164 Tertiary, 41
Spatial heterogeneity, 7, 83 Texas, 190
Spatial variation in fire activity, 6 Thermal damage, 17
Speciation, 31, 54, 57 Thermal imaging, 19
Species replacement, 133, 134, 180, 183 Thermal scarification, 61
Spectral response, 22 Therophytes, 96
Spotting, 20, 96, 109, 110, 159, 166, 169, 187 Thinning, 16, 127, 128, 139, 163, 165, 166
Stand-replacing fires, 63, 69 Topkill, 181, 182
State transition, 133, 135–139, 146, 147 Topographic wind, 5
Stem capacitors, 143 Topography, 17, 79, 159, 167
Stem char, 20 Torching, 17, 109
Stomatal leakiness, 81 Trade-off, 6, 55, 69, 102
Stomatal regulation, 81 Trait composition, 31
Storage, 36, 64, 65, 81, 96, 103, 123, 125, 126, Trait evolution, 32, 54, 56, 57, 68
137, 144, 162 Traits, 2, 3, 8, 10, 11, 26, 31, 37, 53–55, 57,
Stress, 18, 46, 54, 59, 60, 62, 64, 68, 69, 81, 59, 60, 62, 64–69, 76, 82, 94–98, 103,
97, 104, 109, 123, 126, 137, 138, 142, 107, 111, 123, 137, 138, 142, 144, 146,
161–165, 171, 179, 184, 191 148, 158, 159, 161–163, 166–169, 171,
Sub-Saharan Africa, 46 172, 185
Subtropical, 85, 179–182 Transhumance, 45
Subtropical grasslands, 22, 181 Tree diversity, 101
208 Index

Triassic, 40 Water potential, 80, 81


Tropical, 6, 7, 24, 27, 35, 36, 40, 125, 148, Water quality, 37, 38
179–182, 191 Water use, 98, 163, 164, 170, 171
Tropical forests, 35, 36, 45, 179, 180 Water vapor, 35, 78
Turgor, 81 Water yield, 37
Weathering, 33
Wet periods, 16
U Wildfire science, 1, 3
UAVs, 83 Wildfire spread, 10
Understory, 16, 17, 21, 40, 125, 127, 135, 136, Wildland fuel, 1, 2, 11, 33, 75, 76, 93–97, 99,
138, 146, 147, 163–165, 190 100, 102–111, 156, 159
Urbanization, 9, 46, 186–187 Wildland urban interface (WUI), 9, 157,
186, 187
Wind, 5, 17, 77, 79, 110, 159, 167, 169
V Windows of opportunity, 141, 142
Vapor pressure deficit, 5, 76, 78, 86, 123 Winter fires, 24
Vegetation dynamics, 10, 193 Woodland, 35, 43, 79, 84, 181, 185
Vertical fuel continuity, 17, 157 Woody plant encroachment, 181, 182
Volatilization, 8, 38 Woody thickening, 180–181

W X
Warming, 63, 170, 185 Xeriscence, 59, 60
Water cycle = hydrologic cycle, 31, 34, 37, 38 Xylem, 64, 66, 69, 99, 108, 120, 122, 123,
Water points, 157 127, 137, 142–144, 168

You might also like