0% found this document useful (0 votes)
20 views165 pages

smb1de1

Uploaded by

ral.14031981
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
20 views165 pages

smb1de1

Uploaded by

ral.14031981
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 165

SOIL CARBON TEMPORAL DYNAMICS FOLLOWING AGRICULTURAL LAND

ABANDONMENT FROM FIELD TO CONTINENTAL SCALES

Stephen M. Bell

ADVERTIMENT. Lʼaccés als continguts dʼaquesta tesi doctoral i la seva utilització ha de respectar els drets de la
persona autora. Pot ser utilitzada per a consulta o estudi personal, així com en activitats o materials dʼinvestigació i
docència en els termes establerts a lʼart. 32 del Text Refós de la Llei de Propietat Intel·lectual (RDL 1/1996). Per altres
utilitzacions es requereix lʼautorització prèvia i expressa de la persona autora. En qualsevol cas, en la utilització dels
seus continguts caldrà indicar de forma clara el nom i cognoms de la persona autora i el títol de la tesi doctoral. No
sʼautoritza la seva reproducció o altres formes dʼexplotació efectuades amb finalitats de lucre ni la seva comunicació
pública des dʼun lloc aliè al servei TDX. Tampoc sʼautoritza la presentació del seu contingut en una finestra o marc aliè
a TDX (framing). Aquesta reserva de drets afecta tant als continguts de la tesi com als seus resums i índexs.

ADVERTENCIA. El acceso a los contenidos de esta tesis doctoral y su utilización debe respetar los derechos de la
persona autora. Puede ser utilizada para consulta o estudio personal, así como en actividades o materiales de
investigación y docencia en los términos establecidos en el art. 32 del Texto Refundido de la Ley de Propiedad
Intelectual (RDL 1/1996). Para otros usos se requiere la autorización previa y expresa de la persona autora. En
cualquier caso, en la utilización de sus contenidos se deberá indicar de forma clara el nombre y apellidos de la persona
autora y el título de la tesis doctoral. No se autoriza su reproducción u otras formas de explotación efectuadas con fines
lucrativos ni su comunicación pública desde un sitio ajeno al servicio TDR. Tampoco se autoriza la presentación de
su contenido en una ventana o marco ajeno a TDR (framing). Esta reserva de derechos afecta tanto al contenido de
la tesis como a sus resúmenes e índices.

WARNING. The access to the contents of this doctoral thesis and its use must respect the rights of the author. It can
be used for reference or private study, as well as research and learning activities or materials in the terms established
by the 32nd article of the Spanish Consolidated Copyright Act (RDL 1/1996). Express and previous authorization of the
author is required for any other uses. In any case, when using its content, full name of the author and title of the thesis
must be clearly indicated. Reproduction or other forms of for profit use or public communication from outside TDX
service is not allowed. Presentation of its content in a window or frame external to TDX (framing) is not authorized either.
These rights affect both the content of the thesis and its abstracts and indexes.
Cover art: Victor O. Leshyk
SOIL CARBON TEMPORAL DYNAMICS FOLLOWING AGRICULTURAL LAND
ABANDONMENT FROM FIELD TO CONTINENTAL SCALES

Thesis Submitted For Partial Fulfillment of the Requirements for the Degree of:
Doctor of Philosophy

Stephen M. Bell

Supervisors:

Dr. Carles Barriocanal

Dr. César Terrer

Doctoral degree in Environmental Science and Technology


Institute of Environmental Science and Technology
Universitat Autònoma de Barcelona

Bellaterra, June 2022


“On this sand farm in Wisconsin, first worn out and then
abandoned by our bigger and better society, we try to rebuild,
with shovel and axe, what we are losing elsewhere.”
– Aldo Leopold, in the Foreword to
A Sand County Almanac.
Table of Contents
Overview ................................................................................................................................ 1
List of Papers.......................................................................................................................... 3
List of Figures ........................................................................................................................ 4
List of Tables.......................................................................................................................... 6
Acknowledgements ................................................................................................................ 7
CHAPTER I: Introduction ......................................................................................................... 9
1.1 The critical land and soil nexus ................................................................................. 10
1.2 Agricultural land abandonment as a global land use change .................................... 12
1.3 Soil carbon sequestration for climate change mitigation .......................................... 17
1.4 Abandoned agricultural lands as carbon sinks .......................................................... 22
1.5 Statement of the problem .......................................................................................... 30
1.6 Aim & research questions ......................................................................................... 32
1.7 References ................................................................................................................. 33
1.8 Supplementary materials ........................................................................................... 43
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural
land abandonment .................................................................................................................... 45
2.1 Overview ................................................................................................................... 46
2.2 Introduction ............................................................................................................... 47
2.3 Methods ..................................................................................................................... 48
2.4 Results & Discussion ................................................................................................ 48
2.4.1 Proposed management strategies ....................................................................... 48
2.4.1.1 Naturalness ................................................................................................. 50
2.4.1.2 Multifunctionality ....................................................................................... 51
2.4.1.3 Productivity ................................................................................................ 52
2.4.2 Soil carbon sequestration rates of proposals ...................................................... 56
2.5 Conclusions ............................................................................................................... 59
2.6 References ................................................................................................................. 60
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned
agricultural lands ...................................................................................................................... 67
3.1 Overview ................................................................................................................... 68
3.2 Introduction ............................................................................................................... 69
3.3 Methods ..................................................................................................................... 70
3.3.1 Field sites ........................................................................................................... 70
3.3.1.1 Site selection and sampling ........................................................................ 72
3.3.1.2 Sample analysis .......................................................................................... 74
3.3.2 Published data synthesis .................................................................................... 74
3.3.3 Statistical analysis .............................................................................................. 76
3.4 Results ....................................................................................................................... 77
3.4.1 SOC and N concentrations by successional stage and depth ............................. 77
3.4.2 SOC and N stocks by successional stage and depth .......................................... 80
3.4.3 Soil C:N ratios by successional stage ................................................................ 82
3.4.4 Synthesis of chronosequences in Spain ............................................................. 82
3.5 Discussion ................................................................................................................. 84
3.5.1 Post-agricultural SOC and N changes with succession and depth ..................... 84
3.5.2 Post-agricultural SOC changes in Spain ............................................................ 86
3.5.3 Managing post-agricultural SOC accumulation under a changing climate ....... 89
3.6 Conclusions ............................................................................................................... 90
3.7 References ................................................................................................................. 91
3.8 Supplementary materials ........................................................................................... 98
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land
abandonment in Europe ......................................................................................................... 105
4.1 Overview ................................................................................................................. 106
4.2 Introduction ............................................................................................................. 107
4.3 Methods ................................................................................................................... 109
4.3.1 Literature search and data collection ............................................................... 109
4.3.2 Data processing and analysis ........................................................................... 113
4.4 Results & Discussion .............................................................................................. 117
4.4.1 Overall factors driving soil carbon sequestration following ALA................... 117
4.4.2 Overall SOC dynamics following ALA ........................................................... 120
4.4.2.1 Climatic and biogeographical factors on SOC dynamics ......................... 125
4.4.2.2 Land use and management factors on SOC dynamics ............................. 128
4.5 Conclusions ............................................................................................................. 132
4.6 References ............................................................................................................... 133
4.7 Supplementary materials ......................................................................................... 139
CHAPTER V: Conclusions and future perspectives ............................................................. 149
5.1 Conclusions ............................................................................................................. 150
5.2 Future perspectives .................................................................................................. 154
5.3 References ............................................................................................................... 157
Overview
Millennia of intensive and extensive agricultural land use practices has severely depleted global
soil organic carbon (SOC) stocks and drastically increased atmospheric CO2 concentrations.
Yet, for all the billions of hectares of terrestrial ecosystems converted to agriculture, there has
also been hundreds of millions of hectares of agricultural land abandoned. This often-neglected
land use change represents a significant opportunity for ecosystem restoration and soil carbon
sequestration (SCS), especially if ecological succession occurs spontaneously. However,
despite the potential these lands represent for climate change mitigation, they are still
underrepresented in global change science and policy. The impacts of agricultural land
abandonment (ALA) on soil carbon stocks are insufficiently calibrated in biogeochemical
models. We struggle to predict which kinds of agricultural land will sequester, lose, or maintain
pre-existing SOC levels following ALA, and how these trends behave at different spatial scales.
This is especially true in Europe where there has not been a continental-scale synthesis, despite
the widespread extent of past and ongoing ALA and the strong policy interest to increase
European soil carbon stocks.
In this doctoral thesis, I generate new knowledge on the effects of ALA on soil carbon temporal
dynamics, thereby contributing practical information for sustainable land management
decisions involving the land carbon sink. Six major categories of proposed management
strategies for abandoned agricultural lands were identified following a literature review, each
with positive, negative, direct and indirect outcomes depending on site-specific factors and
management objectives. Accordingly, no single strategy is ideal in all scenarios and a
combination of strategies addresses multiple rural development goals concurrently. Focusing
on two of the six strategies (active and passive restoration), I sampled new chronosequences of
ALA to explore the effects of depth and time on SOC at the field scale and I synthesized a new
dataset from peninsular Spain to identify the potential factors responsible for the high
variability in post-agricultural SCS rates observed in the Mediterranean region.
Chronosequence field studies indicate a highly variable process, depending on multiple
environmental and land management factors. The highest rates of SOC accumulation post-
abandonment in Spain can be expected on lands previously used for woody crop production
featuring ~13–17 ° C MAT and ~450–900 mm MAP, with the lowest rates expected on lands
previously used for annual crop production outside this climatic window. From these insights,
I expanded the analysis to the continental scale and assembled the largest dataset ever collected
on SOC stock changes following ALA at known times (n = 804) to investigate the potential
environmental and human management factors driving SCS rates following ALA in Europe.
There is a slow, but significant, rate of SOC stock increase of 1.28% yr–1 (0.32 Mg C ha–1 yr–
1
) on abandoned agricultural lands across Europe. SOC responses were negatively correlated
with initial stock, indicating a soil carbon saturation effect. Abandoned agricultural lands in
biogeographical regions featuring optimal climatic windows had higher SCS rates, but human
management factors can generate both positive and negative effects on SOC, resulting in
several strongly divergent responses to ALA. Past croplands had a notably greater rate of SOC
increase over time than sites that were previously used as pastures, likely a result of lower
initial SOC stocks in croplands compared to pastures. Sites that underwent natural ecological
succession exhibited a greater rate of change in SOC stock compared to sites that were actively
restored or converted to new vegetation land covers, for example through tree planting
practices. These findings suggest that abandoned croplands with low initial SOC stock and

1
managed through natural succession would show the greatest SOC accrual in Europe, while
fertile pastures that are actively converted (e.g., afforested) would result in the lowest increases
in SOC, or even losses.
This work helps clarify some of the previous regional debates on the positive, negative, and
neutral SCS potentials of post-agricultural soils, which have likely been confounded by the
multiple factors identified. Overall, this PhD thesis informs ecosystem restoration policies and
land management strategies on the potential soil carbon benefits, costs, and challenges of post-
agricultural landscapes. The variability in SOC dynamics following agricultural land
abandonment/conversion must be considered in sustainable land use planning that strives to
incorporate the positive ecological and climate change mitigation implications of ALA, taking
into account site-specific conditions and past and present land management factors to avoid
negative impacts for soil health and lost opportunities for climate change mitigation.

2
List of Papers
I. Bell, S.M., Barriocanal, C., Terrer, C. and Rosell-Melé, A., 2020. Management
opportunities for soil carbon sequestration following agricultural land
abandonment. Environmental Science & Policy, 108, pp.104-111.
II. Bell, S.M., Terrer, C., Barriocanal, C., Jackson, R.B. and Rosell-Melé, A., 2021.
Soil organic carbon accumulation rates on Mediterranean abandoned agricultural
lands. Science of the Total Environment, 759, p.143535.
III. Bell, S.M., Terrer, C., Barriocanal, C., Perpiñá Castillo, C., Jackson, R., Franklin,
O, Schillaci, C, Saia, S., Rosell-Melé, A. Factors driving soil carbon sequestration
following agricultural land abandonment in Europe. Manuscript in preparation.
IV. Bell, S.M., Raymond, S.J., Yin, H., Jiao, W., Leshyk, V., Olivetti, E., Terrer, C.
Recarbonizing post-agricultural landscapes. Invited “Comment” manuscript in
preparation for Nature Communications.

3
List of Figures
Figure 1. All spheres in Earth system science are present in soils .......................................... 10
Figure 2. Annual carbon emissions and their partitioning as a function of time ..................... 11
Figure 3. Global spatial estimates of abandoned croplands ..................................................... 13
Figure 4. Categories of marginal lands compared to abandoned agricultural lands ................ 15
Figure 5. Summary of the consequences of ALA .................................................................... 17
Figure 6. The terrestrial carbon cycle featuring natural fluxes and human emissions ............ 18
Figure 7. Pathways of SOC transport, stabilization, and loss .................................................. 21
Figure 8. Landscape progression from frorest, to agriculture, to secondary forest ................. 25
Figure 9. Major categories of management strategies for abandoned agricultural lands ........ 49
Figure 10. Soil carbon sequestration rates of management strategies ..................................... 56
Figure 11. Orthophoto comparison of the Font-rubí chronosequence ..................................... 73
Figure 12. Locations of the published chronosequences and paired plots in Spain ................ 76
Figure 13. Effect of soil depth and successional stage on SOC and N .................................... 79
Figure 14. Change in SOC and N stock (Mg ha–1) for each chronosequence ......................... 81
Figure 15. C:N ratio of different chronosequence stages ........................................................ 82
Figure 16. Relative changes in SOC concentration with time since abandonment for all sites
by past crop type, MAP, and MAT .......................................................................................... 83
Figure 17. Notched boxplots for past crop type, MAP, and MAT ........................................ 100
Figure 18. Distribution of data-pairs by year of publication of the original studies.............. 112
Figure 19. Map of the sampling sites within the extent of the Biogeographical Regions of
Europe .................................................................................................................................... 113
Figure 20. Funnel plot of the full dataset ............................................................................... 116
Figure 21. Effects of dominant non-human-related factors on soil carbon change following
abandonment .......................................................................................................................... 118
Figure 22. Effects of dominant human-related factors on soil carbon change following
abandonment .......................................................................................................................... 119
Figure 23. Relative and absolute change in SOC stock over time since abandonment ......... 122
Figure 24. Relative and absolute change in SOC stock over time since abandonment for
different initial SOC stocks and depths ................................................................................. 124
Figure 25. Relative and absolute change in SOC stock over time since abandonment for MAP
and MAT ................................................................................................................................ 126
Figure 26. Relative and absolute change in SOC stock over time since abandonment for the
Biogeographical Regions in Europe ...................................................................................... 127

4
Figure 27. Relative and absolute change in SOC stock over time since abandonment for past
management type, crop type, and restoration management type ........................................... 130
Figure 28. Model-averaged importance of terms following multi-model analysis ............... 139
Figure 29. Relative and absolute change in SOC stock over time since abandonment for the
full European dataset of ALA ................................................................................................ 152
Figure 30. Global distribution of chronosequences and paired-plot data-pairs ..................... 156

5
List of Tables
Table 1. Summary of large geographic scale meta-analyses of SCS and LULCCs related to
ALA ......................................................................................................................................... 28
Table 2. Typically used delineation of SOM fractions and SOC pools according to the
previous SOM conceptual paradigm ........................................................................................ 43
Table 3. Summary of proposed management strategies and potential new land uses for
abandoned agricultural lands globally ..................................................................................... 54
Table 4. Geographical and soil classifications of the chronosequence field sites ................... 71
Table 5. Soil chemical characteristics of the field sites ........................................................... 78
Table 6. Mean soil 13C and 15N stable isotope values of the field sites ................................ 98
Table 7. Mean soil chemical characteristics of the field sites by depth ................................... 99
Table 8. List of papers included in the Spanish synthesis ..................................................... 101
Table 9. Results of the first stage of data collection process ................................................. 110
Table 10. Results from Kruskal–Wallis tests of significance for factor effects on SOC stock
change .................................................................................................................................... 139
Table 11. Summary of the best-fitting linear model following multi-model analysis ........... 140
Table 12. List of papers included in the European synthesis ................................................. 141

6
Acknowledgements
There are many people I owe thanks for supporting me through the entire PhD, and beyond.
I am extremely lucky to have had such great supervisors: Toni, Carles, and César. I’ve always
felt I had all the bases covered between the three of them to make it through any challenge. I
want to thank Carles and César for supporting me not only in scientific and professional ways,
but in all manner of personal, logistical, and administrative hurdles that came up (and boy,
there were many). As role models and friends, I couldn’t ask for better.
The most important positive influence in my research career has undoubtably been my mentor
and friend, Dr. Antoni Rosell-Melé. From our first meeting (downtown Barcelona after he
performed a castell and introduced me to his family), to our first lab and field work (digging
soil pits and getting lunch at traditional Catalan farmers’ restaurants), to his leadership and
guidance keeping our lab group active, positive, and resolute through a global pandemic, Toni
has inspired me in all aspects of academia, work, and personal life. I owe him enormous thanks
for making my PhD and Barcelona experience so enriching. I dedicate this thesis to his
memory.
To my lab colleagues, old and new, thank you for all the help on science questions, lab
procedures, and everything else. Thank you Pau, Núria M., Mélanie, Maria, Elena, Adrià,
Felipe, Vero, Martí, Gorka, Alex, Joan, Oriol, Núria P., Ashley, Matthieu, Nina, and all other
IMPACTANT and ICTA lab mates. Thank you Helena, Ruofei, Wenzhe, Leila, Maria, Kevin,
Junlan and all other Terrer Lab mates. Thank you to all the ICTA and UAB staff who kept my
degree and visa on track, and allowed me to squeeze in a last-minute overseas research visit
despite a certain university “IT issue”.
Most of all, thanks to my family and loved ones for their unending support during my PhD,
and well, since forever. Thank you, Mom, Dad, Ronnie, Steph, and Matt (and Finley), for
enabling my transatlantic habit, always making sure I’m staying active and healthy, and making
so much effort to spend as much time as possible with me whenever I visit. Despite the great
distance, I’ve never felt far away. And thank you to Maïssane, for keeping me grounded,
positive, and encouraged through all the challenges and stresses a PhD brings. She
(involuntarily) heard more about abandoned soils than anyone else in the world, I’m sure. And,
maybe in solidarity with these abandoned soils, she became an abandoned partner on more
evenings and weekends than I’d like to admit as I worked on this thesis. Thank you for always
being there for me.

7
8
CHAPTER I: Introduction
CHAPTER I: Introduction

1.1 The critical land and soil nexus


In Earth system science, human activity is now conceptualized as the Anthroposphere (Kuhn
and Heckelei, 2010), influencing all other spheres (Figure 1). Unfortunately, the magnitude of
current anthropogenic pressures and their impacts on terrestrial ecosystems is unprecedented:
from higher temperature increases over land compared to oceans via intensified greenhouse
gas emissions (IPCC, 2021), to the exploitation of net primary productivity via biomass
harvesting (Haberl et al., 2014), to the alteration of biogeochemical cycles via chemical inputs
and extractions (Galloway et al., 2008; Lu and Tian, 2017). Human well-being is inextricably
tied to the sustainable management of land and soil resources (Isbell et al., 2017). Natural and
modified landscapes serve as the foundation of human livelihoods through the provisioning of
vital ecosystem services (Hoekstra and Wiedmann, 2014). Therefore, strategically managing
land and soils (i.e., the pedosphere) has become a critical nexus of the Anthropocene (Lewis
and Maslin, 2015), linking human activities with ecosystem stability and climate change (Foley
et al., 2005; Smith et al., 2019; Turner et al., 2007).

Figure 1. All spheres in Earth system science are present in soils (i.e., the pedosphere): soil
air, soil water, soil mineral particles, and soil living and decaying organic matter. Soils are
where the atmosphere, biosphere, hydrosphere, and lithosphere interact. These interactions
are now closely connected to, and sometimes directly modulated by, human activities of the
anthroposphere (e.g., land use and land cover change). Adapted from Lal et al., (1998).
Over three-quarters of Earth’s ice-free land surface is under human management (Erb et al.,
2017) and nearly one-third has experienced land use and land cover change (LULCC) (Winkler
et al., 2021) as a result of either human activities (60%) (Song et al., 2018) or indirect drivers

10
CHAPTER I: Introduction

like climate change. LULCC implies complex interactions and trade-offs to the ecosystem
services soils provide (Smith et al., 2015). Historical and ongoing patterns of LULCC have
stimulated varied soil nutrient and greenhouse gas fluxes (Houghton et al., 2012), with
uncertain implications for regional and global biogeochemical cycling (Peñuelas et al., 2013;
Wieder et al., 2018). The land carbon sink and land carbon flux (Figure 2), for example, is a
function of the interactions and contributions of LULCC, CO2 fertilization, and nitrogen
deposition (Tharammal et al., 2019). However, when the impacts of LULCC on soils are poorly
quantified in Earth system modelling, our ability to predict terrestrial fluxes of the land carbon
sink is significantly weakened (Eglin et al., 2010; Krause et al., 2019; Quesada et al., 2018).
The impacts of LULCC on soil biogeochemical cycling must be constrained to inform
appropriate responses to 21st century land challenges (IPCC, 2019; Smith et al., 2015).

Figure 2. Annual carbon emissions (positive values) and their partitioning (negative values)
as a function of time from the main components of the 2021 Global Carbon Budget
(Friedlingstein et al., 2022).

11
CHAPTER I: Introduction

Soil systems are not static following LULCC (Ryo et al., 2019); the physicochemical legacies
of past disturbances can be measured for decades and longer as soil properties return, often in
a non-linear fashion, to pre-disturbance levels or reach new equilibria (Beniston et al., 2014;
Johannes M. H. Knops and Tilman, 2000; Marin-Spiotta et al., 2009). Despite this, statistically
and geographically robust temporal response curves of soil properties like soil organic carbon
(SOC) to major LULCCs like agricultural expansion and contraction (e.g., abandonment) are
severely lacking in global change science. This lack of theoretical and empirical information
contributes to model inaccuracies, prediction uncertainties, and, ultimately, inadequate land
use policies at a time where informed, sustainable land management is needed more than ever
before (Folberth et al., 2016; Hendriks et al., 2016).

1.2 Agricultural land abandonment as a global land use change


Of all the dominant land uses that significantly impact soils, agriculture (mainly croplands and
pastures) is undoubtably the most pervasive. Approximately 50 million km2 of global soils are
being used for food, feed, fibre, and livestock production (Goldewijk et al., 2011), with humans
claiming 38% of the world’s land area for farmland and appropriating nearly 30% of global net
primary productivity (Haberl et al., 2007; Ramankutty et al., 2008). However, while agriculture
has steadily expanded to every corner of the globe since its advent millennia ago, the reverse
process of agricultural land abandonment (ALA) has been simultaneously occurring. Even with
a 9% increase in total cropland area over the last two decades, there was still 115.5±24.1 Mha
of previously existing cropland that underwent abandonment or conversion (Potapov et al.,
2022). Leirpoll et al., (2021) identified 83 Mha of abandoned cropland from 1992 to 2015. One
of the most often cited global estimates found that between 385–472 Mha of croplands and
pastures were abandoned from the years 1700 to 2000 (Campbell et al., 2008), or between a
quarter to a third of global cropland area (Ramankutty et al., 2018).

Most global and regional estimates of the timing and extent of ALA vary widely, with high
uncertainties for several reasons. For example, it is near-impossible to distinguish abandoned
pastures from natural grasslands and short-term fallow fields using commonly employed global
land cover mapping approaches with remote sensing. This is also why global maps of ALA
focus primarily on abandoned croplands (Figure 3), as they are much easier to detect and
monitor. Small plots sizes in heterogenous and diversely cultivated agrarian regions add more
difficulties, even for cropland detection. However, there have been recent advances in
methodologies enabled by higher spatiotemporal resolution imagery and more accessible cloud
computing (Yin et al., 2020, 2018). As methods and computational powers improve, the first

12
CHAPTER I: Introduction

reliable global maps can be expected, offering information on not only the location of ALA,
but also the timing and duration (especially in the case of cyclical recultivation following
ALA). The overall situation for agricultural land extent has been summarized succinctly in a
recent Our World in Data web article by Dr. Hannah Ritchie: while global croplands are indeed
increasing, the reduction in global pastures has finally decoupled agricultural land expansion
from food production, suggesting that “the world has passed peak agricultural land” (Ritchie,
2022).

Figure 3. Global spatial estimates of abandoned croplands. Top: Cropland extent from 2000–
2019, featuring stable croplands and cropland expansion and reduction (i.e., abandoned or
converted) (Potapov et al., 2022). Bottom: Simplified presentation of abandoned cropland
hotspots from 1992–2015. Visualized in Leirpoll et al., (2021) based on the aggregated gridcell
fraction at 1° resolution. Note the post-Soviet states hotspot following the dissolution of the
Soviet Union (1988-1991) (see Lesiv et al., (2018) for 10 arc second resolution map).

13
CHAPTER I: Introduction

Abandoned agricultural lands can be conceptualized as one of several sub-categories under the
umbrella of “marginal lands”, overlapping primarily with degraded lands because low-
productivity is a common driver of abandonment (Figure 4) (Mellor et al., 2021). There is no
universally recognized definition of ALA with an agreed minimum timeframe due to the wide
array of differing sociocultural and economic perspectives of abandonment as a land use
change. Complication things further, alternative terminologies are also used interchangeably
with “abandoned agricultural lands” in different contexts (e.g., old fields, post-agrogenic, set-
aside, retired land, etc.) and abandonment is by no means a permanent nor one-off change
(Prishchepov et al., 2021). ALA can be cyclical, with periods of abandonment followed by
recultivation followed by abandonment, resulting in conceptual overlaps with fallow lands and
shifting agriculture (Heinimann et al., 2017; Sarkar et al., 2015).

The various definitions of ALA used by land managers and stakeholders, policy makers, and
researchers typically fall under five main categories: administrative, economic, social,
ecological, and agronomic (Anguiano et al., 2008). Regardless of the different strengths,
weaknesses, and specific targets of each category of definitions, they all follow the same basic
tenant: the cessation of agricultural activities and the withdrawal of agricultural management
from the land (Fayet et al., 2022). The Food and Agriculture Organization (FAO) of the United
Nations defines land abandonment as “a process, whereby human control over land (e.g.,
agriculture, forestry) is given up and the land is left to nature. After a number of years,
depending on the ecological zones and climate, land can be considered as completely
“abandoned”, when either legal (e.g., forest law) or natural conditions (e.g., desertification,
overgrowth with forest) render a restoration for agricultural use is impossible or too costly.”
(FAO, 2006). This “number of years” is now generally considered to be at least four
(Prishchepov et al., 2021).

14
CHAPTER I: Introduction

Figure 4. Common categories of marginal lands often compared to abandoned agricultural


lands in the context of soil productivity and ecosystem restoration science (Mellor et al., 2021).
One of the main distinctions of ALA as a global land use change that makes it difficult to map,
monitor, and manage, is the complexity and diversity of potential drivers and impacts. For
example, ALA can be caused by one or more of a range of factors in different regions of the
world (Fayet et al., 2022; Li and Li, 2017; Rey Benayas et al., 2007; Ustaoglu and Collier,
2018). Subedi et al., (2022) identified seven generic categories of drivers (demographic,
household characteristics, farm characteristics, biophysical, economic, regulatory, and socio-
political), and found that some specific drivers were present in all case studies of ALA reported
in the literature: slope, soil quality, land suitability, accessibility of farm and remoteness, off-
farm employment and farm income, migration and depopulation, and farmer age. Ecological
and biophysical drivers include the physiographic and biological factors of the ecosystem when
perceived as constraints to agricultural production, including climate change, whereas land
mismanagement drivers, such as soil degradation due to over exploitation, represent the result
of unsustainable agricultural production over time in a specific ecosystem. Social, economic,
political, demographic, and institutional factors leading to ALA encompass the wide range of
external forces affecting farming profitability, such as market incentives to abandon, migration
to cities and rural depopulation, agricultural industrialization, farmer age and replacement, field
accessibility, proximity to markets, etc. Conflicts and other geopolitical drivers are also an
easily identifiable cause of ALA (Yin et al., 2019). Among many smaller scale national policy
incentives that have been implemented in the last century to abandon croplands, notable large-

15
CHAPTER I: Introduction

scale policies or geopolitical events resulting in widespread abandonment include the breakup
of the Soviet Union (Kuemmerle et al., 2011; Lesiv et al., 2018; Schierhorn et al., 2013) and
the implementation of China’s 1999 Conversion of Cropland to Forest Program, also known as
the “Grain-for-Green” program (Deng et al., 2014a; Gutiérrez Rodríguez et al., 2015).

The impacts of ALA are as diverse and context-dependent as the drivers. Depending on the
perspectives of the observer or stakeholder, ALA can have both positive and negative impacts.
In many cases it is perceived positively as an opportunity for ecosystem regeneration and
climate co-benefits (Navarro and Pereira, 2012; Poore, 2016; Yang et al., 2020), while for
others it signals economic depression, the loss of traditional rural livelihoods, loss of
biodiversity, and an increased risk of wildfires and their effects (Benjamin et al., 2008;
Katayama et al., 2015; Lucas-Borja et al., 2018; Queiroz et al., 2014). The five main negative
impacts identified by Rey Benayas et al., (2007) are: reduction of landscape heterogeneity and
promotion of vegetation homogenisation, soil erosion and desertification, reduction of water
stocks, biodiversity loss and reduced populations of adapted species (compared to biodiverse
agroecosystems) and, lastly, a loss of cultural and aesthetic values. Ustaoglu & Collier (2018)
lists the positive impacts as follows: natural habitat restoration (i.e., rewilding), improvement
in hydrological regulation, decrease in soil erosion, and increases in water quality, soil carbon,
soil fertility, biodiversity, and renewable energy potential. Subedi et al., (2022) surveyed the
consequences of ALA reported in 65 studies from around the world and found further evidence
of these sometimes contradicting positive, negative, and mixed impacts (Figure 5). But from
the perspective of soil functions and ecosystem carbon sequestration at the global scale, ALA
was found to be an almost entirely positive LULCC.

16
CHAPTER I: Introduction

Figure 5. Summary of the consequences of ALA based on a survey of 65 studies from around
the world by Subedi et al., (2022). Note the largely positive impacts of ALA on soils and carbon
sequestration.
Indeed, land use determines the environmental services provided by terrestrial ecosystems,
including the ability to sequester carbon and maintain biodiversity. The return of native
vegetation with increased biomass and soil carbon storage makes ecosystems resilient to
perturbation and supports climate change mitigation (Vilà-Cabrera et al., 2017). Unless the
agricultural land is degraded beyond natural recovery, ALA typical initiates spontaneous
ecosystem regeneration through ecological succession (Cramer et al., 2008; Yang et al., 2020).
Because of the large spatial extent of present and historical ALA and the fact that post-
agricultural soils are often severely carbon depleted, the potential of abandoned agricultural
lands for significant rates of soil carbon sequestration is an important area of research in global
change science.

1.3 Soil carbon sequestration for climate change mitigation


There is more carbon stored in the world’s soils than in the atmosphere and all vegetation
combined (Figure 6). It is the largest reactive pool of carbon in terrestrial ecosystems and may
reach up to 4000 Gt (at a depth of 3 m) when permafrost is fully accounted for (Lal, 2013).
Consequently, because of its large size, potential long-term residence time, and ability to be
manipulated by human management, soil carbon can play a crucial role in balancing the global

17
CHAPTER I: Introduction

carbon budget. Carbon sequestration is the transfer of atmospheric CO2 into longer-term pools
that keep carbon stored in a more stable state, delaying reemission. Soil carbon sequestration
(SCS) is simply this process in relation to the soil component (~2500 Gt) of the global carbon
cycle, increasing both organic and inorganic soil carbon typically through specific land use
management practices. The SOC pool (~1550 Gt) responds more rapidly to human
management than the soil inorganic carbon pool (~950 Gt) and is therefore the main target of
sequestration measures employed in sustainable agriculture and ecosystem restoration
activities (Lal, 2004a). Most agroecosystems have lower than baseline SOC stocks, due to the
long-term negative impacts (e.g., erosion, compaction, mineralization, leaching, etc.) of
agriculture-related processes like ploughing, residue removal, monocropping, and other
intensive farming practices (Lal, 2013; Sanderman et al., 2017).

Figure 6. The terrestrial carbon cycle featuring natural fluxes (green) and human emissions
(red) in Gt of carbon per year. Soils contain more carbon than the atmosphere and vegetation
combined. Numbers in parenthesis represent the estimated size of the carbon pool in Gt
(Trivedi et al., 2018).
Rebuilding carbon stocks in depleted agricultural soils is an effective way to promote food
security and climate change mitigation. The importance of SCS through various sustainable
agricultural practices is recognized globally as a key priority area for the coming years (Bossio

18
CHAPTER I: Introduction

et al., 2020; Bradford et al., 2019; Cornelia Rumpel et al., 2018; Vermeulen et al., 2019), and
it has become increasingly necessary to accurately describe the underlying mechanisms and
nature of soil organic matter (SOM) and SOC stabilization. Not only is SOM fundamental to
soil health and fertility, it is also the largest actively cycling terrestrial carbon pool. It is the
fraction of soil that is composed of living and dead organic materials and residues (i.e., plant
and animal) at varying stages of decomposition representing sources of macro- and
micronutrients for plant growth. Making up approximately only one tenth of total SOM, the
living component consists of live roots and soil organisms generally classified as either
macrofauna (e.g., earthworms, insects, spiders), mesofauna (e.g., springtails, diplura,
enchytraeids), microfauna (e.g., protozoa, nematodes) or microbioata (bacteria and fungi)
(Briones, 2014).

Within SOM, SOC refers only to the carbon component of organic compounds. Because SOM
is difficult to measure directly, researchers prefer to measure, report, monitor, and verify SOC
either through destructive, intensive, and expensive laboratory methods (e.g., physical
fractionation, loss on ignition method, elemental analysis) or through proximal (e.g., VNIR-
SWIR soil spectroscopy) and satellite remote sensing (Angelopoulou et al., 2020; Biney et al.,
2022a, 2022b). Approximately 55–60% of SOM is soil C by mass according to the
conventional understanding, and only one of either of the two properties needs to be measured
in order to infer the other using the “Van Bemmelen conversion factor” (1.724 for determining
soil humus content from a SOC concentration following elemental analysis). This factor has
been used since the late 19th century for converting primary soil data from field sites and for
standardizing large datasets comprising both SOC and SOM values (Minasny et al., 2020).
Although modern review studies argue that a factor of closer to 2 is almost always more
accurate (Pribyl, 2010), the original conversion factor is still regularly used to achieve
consistency throughout datasets (Deng et al., 2016; Don et al., 2011; Kämpf et al., 2016; Liu
et al., 2018).

Starting in the 1970–80’s (see, for example Kögel‐Knabner et al., (1988)), SOM was often
characterized as having four distinct fractions based on size, residence time and chemical
composition: dissolved organic matter; particulate organic matter (POM, comprising of fresh
residues and living components); humus; and resistant organic matter. The classification of
SOM and SOC into fast, intermediate, and slow cycling pools, however, was more of a
conceptual exercise, being operationally defined rather than a measurable attribute (see
Supplementary materials Table 2 for an outline of the previously dominant SOM/SOC

19
CHAPTER I: Introduction

framework) (Lehmann and Kleber, 2015). Over subsequent decades and especially around the
turn of the millennium, SOM researchers began to demonstrate that the compounds previously
considered recalcitrant where in fact degradable under the right conditions (Gleixner et al.,
2002; Rasse et al., 2006). New pools were identified largely based on the percentage of SOC
that is environmentally susceptible to microbial activities (Schmidt et al., 2011), physically
protected in aggregates (Tisdall and Oades, 1982), or mineral-associated organic matter
(MAOM) following standard analytical techniques (Torn et al., 1997). Indeed, both SOM and
SOC are now described as continuums of organic material mixtures undergoing constant back-
and-forth transformation between decomposition and stabilization (Lehmann and Kleber,
2015), best defined by using mainly the POM and MAOM fractions (Lavallee et al., 2020).

Unfortunately, some of the traditional conceptions of SOM processes, which do not represent
the diverse conditions of soils found across the globe, still inform many biogeochemical
models. For example, model results might predict greater SOM storage and slower turnover
for finely textured soils, by generally assuming high clay and silt content are good indicators
of SOM-stabilizing conditions (i.e., greater aggregation, sorption, soil moisture, etc.) and by
describing SOM storage based on the presence of arbitrary carbon pools with varying turnover
times (Rasmussen et al., 2018). But in reality, recent findings indicate that all clay-sized
particles may not have the same effect on SOM storage, and other physicochemical attributes,
such as sorption and the content of extractable metals, might be more indicative of SOM
stabilization potential. Because sorption (or the formation of chemical associations between
minerals and organic compounds) can aid in the stabilization and protection of even labile or
young compounds (Abramoff et al., 2021), it is now generally considered one of the key
mechanisms building MAOM as a carbon pool and it is being increasingly adopted in SOC
models (Figure 7) (Schmidt et al., 2011; Sulman et al., 2018). SOM stabilization enables long-
term SCS, and according to the new paradigm recently championed by Lehmann and Kleber
(2015), there are two primary mechanisms driving this process: the formation of MAOM and
the formation of soil aggregates which protect SOM (Angst et al., 2021), both which are
promoted following the cessation of destructive agricultural practices.

20
CHAPTER I: Introduction

Figure 7. Pathways of SOC transport, stabilization (i.e., as MAOM), and loss through the
plant-soil interface. Land use determines the quality and quantity of each pathway in human
managed soil systems. Adapted from Dynarski et al., (2020).
Aside from collecting new in-situ soil data from field sites to determine SCS rates (whether
long-term experimental stations or one-time sampling campaigns), modelling approaches can
integrate multiple predictor factors to estimate farm-, regional-, and global-scale sequestration
potentials under various management regimes. In the European Union, several different models
have recently been employed to up-scale field data and project SOC stock fluctuations on active
agricultural lands into the future based on potential EU and individual member state policy
scenarios, including MITERRA-NL in the Netherlands, C-TOOL in Denmark, CENTURY and
RothC in Spain, RothC and RothC10N in Italy, EPIC in Germany and Italy, DNDC in Poland,
ICBM in Sweden, and PaSim and STICS in France (see Rodrigues et al., (2021) for individual
study references).

Maximizing SOC storage is also the best strategy to improve overall soil health, while
increasing the global land carbon sink (Figure 2) (FAO and ITPS, 2021). A healthy soil system
has the capacity to provide multiple ecosystem services, especially soils managed sustainably
with carefully selected practices (Paustian et al., 2016). In recent decades, several sustainable
agricultural approaches that ensure food, feed, and fibre production while promoting SCS and
other soil health measures have been developed as alternatives to intensive conventional
practices. These include popular approaches like agroecology, nature-inclusive agriculture,
permaculture, biodynamic agriculture, organic farming, conservation agriculture, regenerative

21
CHAPTER I: Introduction

agriculture, carbon farming, climate-smart agriculture, high nature value farming, low external
input agriculture, circular agriculture, ecological intensification, and sustainable intensification
(Oberč and Arroyo Schnell, 2020). Within these approaches, the specific sustainable
agricultural practices employed can help promote faster rates of SCS on many of the key land
uses dominating the surface of the Earth, whether croplands, pastures and grazing lands, or
managed forests. At the most extreme, one strategy could simply be the conversion from a low-
SOC land use into a land use with a higher SOC base level (i.e., higher SOC equilibrium). But
other SCS-promoting practices are less disruptive and do not require a complete change in land
use, such as promoting vegetation regimes with higher carbon inputs (e.g., crop rotation, cover
crops, perennial crops, etc.), managing applied nutrient regimes (e.g., optimized fertilizer rate,
type, timing, and precision, etc.), and protecting soil and water properties (e.g., reduced tillage,
no-tillage, crop residue retention, reduced compaction and erosion, improved water/irrigation
regimes, etc.) (Smith et al., 2019). Sustainable agricultural practices imply actions taken with
a holistic understanding for the integrated agroecosystem of soil (e.g., no-tillage, minimum
tillage, hedgerows, erosion and compaction management, etc.), crop (e.g., cover, intercropping,
rotation, residues, legume incorporation, etc.), water (e.g., irrigation regime, harvesting, reuse
practices, etc.), biodiversity (e.g., pollination management, pest management, agro-ecological
measures, etc.), and inputs (e.g., dose size and timing of amendments like biochar, manure,
compost, litter, mulching, etc.).

1.4 Abandoned agricultural lands as carbon sinks


In the global effort to mitigate the environmental and social impacts of climate change, highly
complex solutions are often proposed that involve significant technical and financial
intervention. Recent sentiment in the land sector, however, has been shifting towards more
naturally inspired solutions, commonly known as nature based solutions (Cohen-Shacham et
al., 2016; Keesstra et al., 2018) or natural climate solutions (Bossio et al., 2020; Griscom et
al., 2017) depending on the objectives and approaches. For example, instead of large-scale
afforestation with single species tree plantations to restore ecosystems, holistic ecosystem
regeneration practices in line with natural successional processes has been shown to be a more
effective strategy for achieving carbon sequestration and other overarching environmental
goals (Seddon et al., 2019). Reducing atmospheric CO2 concentrations through sequestration
in the pedosphere and biosphere has many potential advantages. It serves to mitigate climate
change, improve ecosystem resilience, and, when it is achieved through sustainable agricultural
practices, it can increase SOC, improve soil quality, advance global food security and possibly

22
CHAPTER I: Introduction

provide another economic generation stream for farmers (i.e., carbon credits) (FAO and ITPS,
2021; Lal, 2008). As most nations have committed to limiting global average temperature rise
to well below 2° C, soils are now central to global climate change mitigation efforts (Minasny
et al., 2017). This has resulted in the consensus that carbon-depleted agricultural soils have the
greatest potential for immediate action to promote carbon sequestration as a natural climate
solution (Amelung et al., 2020).

Soils as a whole still contain more carbon than plants and the atmosphere combined (Figure 6),
but the depletion of agricultural SOC stocks by intensive practices throughout human history
(12 millennia of agriculture) has left a global carbon debt of approximately 116 Gt (“Correction
for Sanderman et al., Soil carbon debt of 12,000 years of human land use,” 2018; Sanderman
et al., 2017), roughly equivalent to all the CO2 emitted by the United States since 1800. Since
the Industrial Revolution alone, 78±12 Pg C has been released to the atmosphere from soils
used as croplands, pastures, and rangelands, together covering nearly half of the Earth’s surface
that can support vegetation (Bondeau et al., 2007; Foley et al., 2005; Lal, 2004b). Fortunately,
carbon-depleted soils have the ability sequester SOC, either naturally or through specific SOM-
friendly management practices as described in the previous section. The leading international
initiative aiming to leverage the climate mitigation potential of carbon sequestration in
agricultural soils is known as “4 per 1000” Soils for Food Security and Climate. Launched at
UNFCCC COP21 by the French Ministry of Agriculture, it suggests that by increasing global
soil organic matter by 0.4% per year through improved agricultural practices, 2–3 Gt C year–1
could be sequestered, which would effectively offset 20–35% of anthropogenic greenhouse gas
emissions (Minasny et al., 2017). Although it may not be necessary nor productive to target
such a specific and potentially unrealistic sequestration rate (de Vries, 2018; Poulton et al.,
2018; White et al., 2018), there is no debate that SOC on active agricultural lands globally must
be increased and stabilized as much as possible within ecological and socioeconomic limits
(Bossio et al., 2020; Bradford et al., 2019; Vermeulen et al., 2019; Zomer et al., 2017). With
that being said, the cessation of agriculture altogether (i.e., ALA) is often the most efficient
way to restore ecosystems and sequester carbon in tandem and at scale.

Due to the SOC-depleted nature of agricultural lands, the widespread historical and ongoing
prevalence of ALA, and the innate ability of soils to rebuild SOC stocks, abandoned
agricultural lands represent some of the largest human-induced carbon sinks ever measured. In
the post-Soviet states, vast expanses of forests regrew over the 62.6 Mha of croplands
abandoned following the collapse of the Soviet Union (Schierhorn et al., 2019), and in this

23
CHAPTER I: Introduction

process drew enormous quantities of carbon down into the above- and belowground biomass
and soils (Henebry, 2009). Estimates of the size of this sink range in the hundreds of teragrams,
depending on the spatial extent considered, carbon pools included, and methods of analysis
(Kuemmerle et al., 2011; Schierhorn et al., 2013; Vuichard et al., 2008), with significant annual
SOC accrual rates comparable to intentional restoration initiatives (Dymov et al., 2018;
Kalinina et al., 2015; Kurganova et al., 2015, 2014; Wertebach et al., 2017). Further back in
human history, the deadly arrival of Europeans to the Americas tragically lead to the
abandonment of an estimated 55.8 Mha, which subsequently regrew into secondary forests,
sequestering 7.4 Pg C and possibly intensifying the post-medieval Little Ice Age (Koch et al.,
2019). On a more practical level, ambitious regional efforts that intentionally restore active and
abandoned agricultural lands, like the “Grain-for-Green” program in China and the
Conservation Reserve Program in the USA, have also demonstrated the significant ecological
and carbon sink co-benefits possible following agricultural cessation (Deng et al., 2014a,
2014b; Munson et al., 2012; Robles and Burke, 1998; Shi and Han, 2014).

Once agricultural activities are abandoned, there are multiple ways to measure and calculate
the SCS rates present. Time-stamped data points are crucial for understanding what factors
determine if and when a given plot of abandoned land will act as a carbon sink or source
following agricultural cessation. The gold-standard is unquestionably repeated measurements
on the same plot of land over time, ideally in a long-term experimental setting, as it will have
the most reliable and representative conditions. However, global biogeochemical models
including ALA are currently limited by relatively poor temporal SOC data (i.e., low quality
and quantity) due to the logistical and financial challenges of long-term field sites with repeated
measurements, especially in under-resourced regions. Statistically robust and geographically
representative temporal SOC response curves are needed to improve model accuracy and
prediction certainty, and thereby strengthen land management policies that incorporate ALA
and ecosystem restoration objectives for agriculture lands.

Investing new money, time, and energy in long-term field sites may not be necessary nor
practical. Previously published space-for-time substitutions, like paired-plots and
chronosequences (Figure 8), are readily available alternative data sources that can be used to
complement datasets of repeated field measurements because they have not yet been
sufficiently synthesized at different geographic scales (mainly a lack of comprehensive
continental and global syntheses) (Huggett, 1998). Paired-plots of ALA comprise of one
control plot under active agricultural practices (i.e., stage 0) paired with one field plot where

24
CHAPTER I: Introduction

the same agricultural practices have since ceased at a known time in the past (years).
Chronosequences of ALA are simply a series of two or more field plots differing in time since
abandonment, each paired with the same control plot. All plots must be comparable in all other
environmental and human management factors (i.e., similar soil type, climate, vegetation,
management and cropping practices, restoration practices, etc.), such that the modulating
effects of time since abandonment on the investigated variable (e.g., soil biological, chemical,
and physical properties) is isolated. Due to these sampling constraints, the individual field plots
of paired-plots and chronosequences are typically identified and selected in as close proximity
to each other as possible to ensure the similarity of environmental and land use history factors.

Figure 8. Typical landscape progression involving (left-to-right) land clearing from primary
forest, agriculture, agricultural cessation, and ecological succession into secondary forest.
Chronosequences of ALA identified in the field require multiple individual plots: a control plot
under active agricultural practices (i.e., stage 0) and associated abandoned plots at known
times (years) since agricultural practices (i.e., stages 1, 2, 3, etc.). When sampled collectively
and analyzed sequentially, these plots represent the temporal trajectory of SOC and other
parameters during ecological succession in response to historical ALA. Under normal
conditions without stalled succession (i.e., if the soil is not degraded beyond recovery), ALA
initiates the spontaneous regeneration of below- and aboveground biomass (i.e., carbon
stocks) and the return of pre-agricultural soil physicochemical conditions. Traditional theory
presumes that SOC is lost following the clearing of land and the initiation of agricultural
practices, followed by a period of flux immediately after ALA and a period of gradual
accumulation during ecosystem recovery until reaching saturation (i.e., a new soil carbon
equilibrium).
Repeated measurements sample one field site repeatedly, meaning the samplings are separated
through time but not space. As space-for-time substitutions, paired plots and chronosequences
sample multiple field sites (separated in space) at the same time, with time since abandonment
accounted for a priori through knowledge of the land use history. While they are not ideal, they
are logistically superior to repeated measurements and generally informative for exploring
broad ecological theories, especially when investigating processes that can surpass the career
or life span of researchers (e.g., SCS) (Walker et al., 2010). By extracting, synthesizing, and
repurposing time-stamped SOC data from any published study involving the use of paired-
plots or chronosequences of ALA, no matter the original research angle so long as SOC is

25
CHAPTER I: Introduction

reported, we can more robustly benchmark and validate models of successional carbon
dynamics.1 Early limited attempts of this at regional and national scales have already been
performed, but not robustly and not at continental nor global scales, resulting in an incomplete
understanding of one of the world’s major LULCCs: ALA.

To overcome spatial and inferential limitations of individual experiments, and to detect


underlying driving factors for many processes present in ecology, one of the best statistical
approaches is to synthesize response ratios of multiple studies and perform meta-analyses
(Gurevitch et al., 2001; Hedges et al., 1999). In the case of the response of soil carbon to ALA,
there has not been a dedicated, comprehensive, and statistically robust global study.
Nevertheless, there have been two seminal studies exploring SOC responses that include ALA
among other LULCCs, originally published over two decades ago and receiving increasing
interest: Guo and Gifford (2002) and Post and Kwon (2000). As of June 2022 on Google
Scholar, Guo & Gifford (2002) has been cited 4001 times (an increase of 2505 since January
2019), while Post & Kwon (2000) has been cited 3201 times (increase of 2083 since January
2019). Yet, despite the high interest and fact that they are global in scale, both studies are
severely data sparse (e.g., < 50 observations for ALA related LULCCs, such as crop to
grassland or forest conversions). Guo & Gifford (2002) found an overall 53% increase in SOC
stock following crop to secondary forest; however, the authors noted that the low quantity of
available data combined with the diversity of methodologies used prevented strong statistical
conclusions. The authors also noted that broadleaf tree plantations on native forest land or
pastures did not affect SOC stock, while pine plantations reduced C stock. This early finding
in the literature supports the recent shift for more research into natural climate solutions like
restoration of abandoned agricultural lands over tree plantations as SCS strategies. Post &
Kwon (2000) conducted a more in-depth review of SOC dynamics during natural regeneration
after agricultural cessation and found average rates of accumulation to be 33.8 and 33.2 g C m-
1
y-1 for forests and grasslands, respectively. They found this level of sequestration, when
considering the global land area potentially affected by ALA at the time, to be relatively small
in comparison to the estimated rate of total carbon sequestration occurring in the Northern

1
This is also an intentional exercise in repurposing past data, revaluing past investments (e.g., research funding),
and recapturing the often neglected inferential and transformative potential of the “long tail of dark data” (Heidorn,
2008; Novick et al., 2018) (i.e., uncompiled, underrepresented, and unique past research at risk of disappearing
due to the threat of continual data availability loss (Vines et al., 2014)).

26
CHAPTER I: Introduction

Hemisphere in biomass and surface litter. However, as in the case of Guo & Gifford (2002),
this review includes only a small number of observations (47) and uses unreliable estimates of
the extent of land abandonment in the Northern Hemisphere (e.g., all sources are from before
the dissolution of the Soviet Union (1988–1991)).

In the two decades since Guo & Gifford (2002) and Post & Kwon (2000), there have been
hundreds of individual field studies on post-agricultural SOC dynamics, which have since been
more comprehensively summarized in a few synthesis studies at regional and biome scales
(Table 1). Don et al., (2011) reviewed 385 studies on LULCC in the tropics and found that a
conversion of cropland to secondary forest resulted in an increase of SOC by 50.3±11.9% over
an average of 32±7 years, with lesser increases for cropland to grassland (25.7±11.1% over
21±6 years) and cropland to fallow (32.2±16.1% over ≤ 7 years). Their study was restricted to
the tropics and, presumably because the authors considered several other LULCC types, there
were relatively few observations compiled for cropland revegetation conversions (25 for
cropland to secondary forest; 16 for cropland to grassland; 21 for cropland to fallow).
Additionally, the paper specifically refers to afforestation and not natural vegetation succession
nor abandonment. Poeplau et al., (2011) found that for a cropland to grassland LULCC in the
temperate zone, SOC accumulated at a rate of 40±11% over 20 years and 128±23% over 100
years based on the model predictions. This was the highest gain among the LULCCs considered
in their study. For cropland to forest, SOC increased at a rate of 16±7% over 20 years and
83±39% over 100 years. This paper also included a limited number of observations: 89 for
cropland to grassland and 70 for cropland to forest. The authors also specifically use the term
afforestation, so it is unclear how many, if any, of the 15 observations of the cropland to forest
LULCC category were natural vegetation succession (i.e., reforestation), which is the trajectory
most representative of ALA. Kämpf et al., (2016) assessed the potential of temperate
agroecosystems for SCS under different climatic and edaphic conditions. Their study looked at
global temperate soils and identified, among other LULCCs, 54 observations of SCS as a result
of ALA from 17 publications. Over an average of 14 years since abandonment, SOC stocks
increased by 18%, with a sequestration rate of 0.72 t C ha–1 y–1. The authors also found that
SCS may not be limited by low primary productivity within the temperate zone, providing
incentive for more comprehensive analyses at different spatial scales and comparing multiple
biomes. W. Li et al., (2018) compiled a global dataset of 836 observations of grassland-land
related LULCC, including 194 observations of cropland to grassland conversion, and the
subsequent changes in SOC stock. The cropland to grassland LULCC category resulted in a

27
CHAPTER I: Introduction

SCS rate of 4.3 kg C per m2 (median) over 79 years (median), a relative increase of 46%. While
their study offers a useful dataset, it unfortunately does not include the key LULCC of cropland
to forest because the authors specifically targeted only grassland-related changes. Furthermore,
the authors used satellite-based net primary productivity (NPP) observations as a proxy for
carbon input.

Table 1. Summary of large geographic scale meta-analyses of SCS and LULCCs related to
ALA.

Authors Year Region LULCC N. of Observations N. of Studies SCS Years


Guo & Gifford 2002 Global C to F 9 ? 53% ?
Post & Kwon 2000 Global C to F 47 28 33.8g C m-2 y-1 n/a
Laganière, et al. 2010 Global C to F 92 ? 26% ?
Don, et al. 2011 Tropical C to F ? 25 50.3±11.9% 32
Don, et al. 2011 Tropical C to G ? 16 25.7±11.1% 21
Poeplau, et al. 2011 Temperate C to F 70 15 16±7% 20
Poeplau, et al. 2011 Temperate C to F 70 15 83±39% 100
Poeplau, et al. 2011 Temperate C to G 89 24 39.8±11% 20
Poeplau, et al. 2011 Temperate C to G 89 24 128.4±23.2% 100
Kämpf, et al. 2016 Temperate C to G 54 17 18% 14
W Li, et al. 2018 Global C to G 194 ? 46% 79
Deng, et al. 2016 Global C to F 31 ? Not significant n/a
-1 -1
Deng, et al. 2016 Global C to G 57 ? 0.30 Mg ha yr n/a
(C to F: cropland to forest; C to G: cropland to grassland; ? : not indicated)

Elsewhere in the world, individual studies that include temporal SOC data following ALA can
be found on every continent, and national-scale meta-analyses synthesizing SOC dynamics
after ALA and similar post-agriculture LULCCs have been undertaken in China, Russia, and
Northern Europe. Deng, Liu, et al., (2014) synthesized 135 studies, which included 844
observations at 181 sites, to determine the effects of China’s “Grain-for-Green” Program. The
authors found a SCS rate of 0.33 Mg ha–1 yr–1 in the top 20 cm of soil. While Hong et al., (2020)
is one of the most recent and data-rich SOC studies with 619 afforested paired-plots, it only
covers northern China, and mixes post-agricultural soils with other non-forested soils like
barren land, grassland, natural forest and riparian sand land. Kurganova et al., (2014) compiled
a database of 116 paired plots from 45 sites across Russia and calculated an average SCS rate
of 0.96 Mg ha–1 yr–1 in the top 20 cm of soil over the first 20 years since abandonment. In their
meta-analysis of SCS following afforestation in Northern Europe, Bárcena et al., (2014) did
not consider natural forest regrowth, but found the largest increase in SOC occurred on former
croplands (compared to former grasslands, heathlands, and barren lands) at 20% for the 0–10

28
CHAPTER I: Introduction

cm depth based on 51 observations. Studies such as these provide useful databases and
reference lists that can be included into larger continental or global meta-analyses.

Most meta-analyses indicate a positive effect on SOC stocks after conversion of croplands to
forests globally. However, there still remains significant data and knowledge gaps and a need
for more robust and specific continental and global meta-analyses explicitly dedicated to SCS
following ALA, as a globally relevant LULCC. Indeed, ALA has occurred anywhere in the
world where natural land has been converted into agricultural use, and has the potential to
continue to occur as long as environmental, social, and economic factors remain dynamic.
When viewed as a climate change mitigation strategy, it is arguably the only strategy that does
not require any additional resources to initiate and is possible to implement anywhere in the
world when conditions permit (e.g., when conflicts with food production, agrobiodiversity, or
land rights, etc., are not present). It is now clear we must consider integrating ALA and
protecting existing abandoned agricultural lands whenever possible to diversify the global land
carbon sink and support the UN Decade of Ecosystem Restoration (2021-2030) (Abhilash,
2021; Aronson et al., 2020). However, ALA is poorly represented in terrestrial carbon models,
both spatially as a land classification and temporally as carbon sinks. The lack of temporal SOC
data in particular has resulted in severe uncertainties on the intensity, the longevity, and the
modulating factors of SCS following ALA. This hinders our ability to monitor, quantify, and
leverage ALA processes strategically, precisely when we need to most.

29
CHAPTER I: Introduction

1.5 Statement of the problem


1. Demand for policy support: Despite the potential abandoned agricultural lands
represent for climate change mitigation through SCS, they are still underrepresented in
global change science and policy. ALA is a global land use change present in every
agricultural region of the world. Due to its contentious sociocultural aspects, there is no
standard framework on how abandoned agricultural lands should be managed for
ecosystem and climate-related goals. In Europe, not only is ALA expected to be a
significant driver in land use dynamics in the coming decades, but most EU climate
policies dealing with land explicitly call for dedicated attention given to rebuilding the
continent’s soil carbon pool. There is currently a strong need for increased scientific
research efforts to support policy decisions concerning the SCS implications of
processes like ALA.
2. Logistical barriers and data limitations: One of the most important analytical
dimensions needed to be able to fully evaluate the carbon sink potential of ALA at large
geographic scales and over longer policy horizons is still severely lacking: namely,
time. Current state-of-the-art biogeochemical models that incorporate the effects of
ALA on soil carbon stocks are limited by relatively poor temporal data (i.e., low sample
sizes) due to the logistical and financial challenges of long-term field studies with
repeated measurements, especially in under-resourced regions of the world. To produce
robust datasets considering the full temporal dynamics of SOC following ALA,
alternative methodologies that are faster, more accessible, and less expensive are
needed. Fortunately, there exists thousands of published studies with data that provides
the necessary temporal resolution, spatial coverage, and combinations of soil type, crop
type, and management practices. Instead of resource intensive repeated measurements,
these studies utilize chronosequences (i.e., simultaneously sampled soil plots sharing
key environmental and experimental conditions but differing in time since the
implementation of the investigated practice). Chronosequence data has not yet been
comprehensively extracted, standardized, and analyzed to assess the temporal dynamics
of SOC following ALA.
3. Unknown soil carbon sequestration rates following ALA: Insufficient calibration of
temporal soil carbon trends directly results in lost opportunities to prepare for the long-
term impacts of ALA and to properly account for its role as a driver or mitigator of
climate change. SCS rates following abandonment are either statistically weak or non-

30
CHAPTER I: Introduction

existent in many regions of Europe, let alone the world. We lack the ability to accurately
predict whether any given plot of agricultural land will sequester, lose, or maintain pre-
existing SOC levels following abandonment, and we do not know how any of these
SOC trends will behave through time or at different geographic spatial scales.
4. Unknown modulating factors of soil carbon sequestration rates following ALA:
Due to the issues mentioned above, the current conceptual framework explaining the
differences in sequestration rates observed worldwide is underdeveloped. This is
especially true in Europe where there has not been a continental-scale analysis, despite
the widespread extent of past and ongoing ALA and the clear motivation to increase
European soil carbon stocks. Researchers have not reached a theoretical consensus on
the driving factors behind different sequestration rates that may explain why abandoned
croplands may lose SOC in one area of Europe while they gain SOC in other areas.
Site-specific conditions and human management factors have not been
comprehensively explored in the literature; they are still poorly quantified due to the
complexity of possible interactions.

These problems and research gaps hinder decisionmakers and land managers from
promoting (or deterring) ALA in rural areas as a component of regional climate
initiatives—inconsistent with sustainable land management objectives. Guidance on
management pathways based on accurate representation of the temporal responses of SOC
to ALA is a top-priority.

31
CHAPTER I: Introduction

1.6 Aim & research questions


The overall aim of this PhD thesis is to generate new knowledge on the effects of agricultural
land abandonment on soil carbon stocks, thereby contributing practical information for
sustainable land management decisions involving climate change mitigation. I investigate the
temporal dynamics of SOC following the cessation of agricultural activities at the field (i.e.,
Province of Barcelona), regional (i.e., Spain), and continental (i.e., Europe) scales. Ultimately,
this thesis provides novel insights into the capacity of European agricultural soils to
recarbonize through ecological succession, whether natural or assisted. The following research
questions (RQ) are addressed:

RQI. What are the main categories of sustainable land management options for abandoned
agricultural lands proposed by researchers, and how do they impact soil carbon
stocks?
I conduct a literature review to identify the main categories of sustainable land management
strategies that have been proposed for existing abandoned agricultural lands or for active
agricultural lands directly converted. Following this review, I then compare the SCS rates
reported in the published literature for these categories.

RQII. How does soil organic carbon in Spain respond to agricultural land abandonment, and
what are the modulating factors?
I undertake field work in the Province of Barcelona, Spain to explore the effects of depth and
time on SCS following agricultural land abandonment. I then synthesize published
chronosequence data across peninsular Spain to identify the potential factors responsible for
the high variability in post-agricultural SCS rates observed in the Mediterranean region.

RQIII. How do soil organic carbon across Europe respond to agricultural land abandonment,
and what are the modulating factors?
I compile and synthesize published chronosequence data of SOC change over time on post-
agricultural lands in European countries, exploring potential environmental and human
management factors driving SCS. I investigate the conditions by which the highest and lowest
sequestration rates may be expected, and provide new insight into the capacity of European
abandoned agricultural lands to serve as carbon sinks into the future.

32
CHAPTER I: Introduction

1.7 References
1. Abhilash, P. C. (2021). Restoring the Unrestored: Strategies for Restoring Global Land during the UN Decade
on Ecosystem Restoration (UN-DER). Land, 10(2), 201. https://doi.org/10.3390/land10020201

2. Abramoff, R. Z., Georgiou, K., Guenet, B., Torn, M. S., Huang, Y., Zhang, H., Feng, W., Jagadamma, S.,
Kaiser, K., Kothawala, D., Mayes, M. A., & Ciais, P. (2021). How much carbon can be added to soil by
sorption? Biogeochemistry, 152(2–3), 127–142. https://doi.org/10.1007/s10533-021-00759-x

3. Amelung, W., Bossio, D., de Vries, W., Kögel-Knabner, I., Lehmann, J., Amundson, R., Bol, R., Collins, C.,
Lal, R., Leifeld, J., Minasny, B., Pan, G., Paustian, K., Rumpel, C., Sanderman, J., van Groenigen, J. W.,
Mooney, S., van Wesemael, B., Wander, M., & Chabbi, A. (2020). Towards a global-scale soil climate
mitigation strategy. Nature Communications, 11(1), 5427. https://doi.org/10.1038/s41467-020-18887-7

4. Angelopoulou, T., Balafoutis, A., Zalidis, G., & Bochtis, D. (2020). From Laboratory to Proximal Sensing
Spectroscopy for Soil Organic Carbon Estimation—A Review. Sustainability, 12(2), 443.
https://doi.org/10.3390/su12020443

5. Angst, G., Mueller, K. E., Nierop, K. G. J., & Simpson, M. J. (2021). Plant- or microbial-derived? A review
on the molecular composition of stabilized soil organic matter. Soil Biology and Biochemistry, 156, 108189.
https://doi.org/10.1016/j.soilbio.2021.108189

6. Anguiano, E., Bamps, C., Terres, J., Pointereau, P., Coulon, F., Lambotte, M., Stuczynski, T., Sanchez
Ortega, V., & del Rio, A. (2008). Analysis of the driving forces behind farmland abandonment and the extent
and location of agricultural areas that are actually abandoned or are in risk to be abandoned.

7. Aronson, J., Goodwin, N., Orlando, L., Eisenberg, C., & Cross, A. T. (2020). A world of possibilities: six
restoration strategies to support the United Nation’s Decade on Ecosystem Restoration. Restoration Ecology,
28(4), 730–736. https://doi.org/10.1111/rec.13170

8. Bárcena, T. G., Kiær, L. P., Vesterdal, L., Stefánsdóttir, H. M., Gundersen, P., & Sigurdsson, B. D. (2014).
Soil carbon stock change following afforestation in Northern Europe: A meta-analysis. In Global Change
Biology. https://doi.org/10.1111/gcb.12576

9. Beniston, J. W., DuPont, S. T., Glover, J. D., Lal, R., & Dungait, J. A. J. (2014). Soil organic carbon dynamics
75 years after land-use change in perennial grassland and annual wheat agricultural systems.
Biogeochemistry, 120(1–3), 37–49. https://doi.org/10.1007/s10533-014-9980-3

10. Benjamin, K., Bouchard, A., & Domon, G. (2008). Managing abandoned farmland: The need to link
biological and sociological aspects. Environmental Management, 42(4), 603–619.
https://doi.org/10.1007/s00267-008-9176-5

11. Biney, J. K. M., Blöcher, J. R., Bell, S. M., Borůvka, L., & Vašát, R. (2022). Can in situ spectral measurements
under disturbance-reduced environmental conditions help improve soil organic carbon estimation? Science
of The Total Environment, 838, 156304. https://doi.org/10.1016/j.scitotenv.2022.156304

12. Biney, J. K. M., Vašát, R., Bell, S. M., Kebonye, N. M., Klement, A., John, K., & Borůvka, L. (2022).
Prediction of topsoil organic carbon content with Sentinel-2 imagery and spectroscopic measurements under
different conditions using an ensemble model approach with multiple pre-treatment combinations. Soil and
Tillage Research, 220, 105379. https://doi.org/10.1016/j.still.2022.105379

13. Bondeau, A., Smith, P. C., Zaehle, S., Schaphoff, S., Lucht, W., Cramer, W., Gerten, D., Lotze-campen, H.,
Müller, C., Reichstein, M., & Smith, B. (2007). Modelling the role of agriculture for the 20th century global
terrestrial carbon balance. Global Change Biology. https://doi.org/10.1111/j.1365-2486.2006.01305.x

14. Bossio, D. A., Cook-Patton, S. C., Ellis, P. W., Fargione, J., Sanderman, J., Smith, P., Wood, S., Zomer, R.
J., von Unger, M., Emmer, I. M., & Griscom, B. W. (2020). The role of soil carbon in natural climate
solutions. Nature Sustainability. https://doi.org/10.1038/s41893-020-0491-z

15. Bradford, M. A., Carey, C. J., Atwood, L., Bossio, D., Fenichel, E. P., Gennet, S., Fargione, J., Fisher, J. R.
B., Fuller, E., Kane, D. A., Lehmann, J., Oldfield, E. E., Ordway, E. M., Rudek, J., Sanderman, J., & Wood,

33
CHAPTER I: Introduction

S. A. (2019). Soil carbon science for policy and practice. In Nature Sustainability (Vol. 2, Issue 12, pp. 1070–
1072). Nature Research. https://doi.org/10.1038/s41893-019-0431-y

16. Briones, M. J. I. (2014). Soil fauna and soil functions: A jigsaw puzzle. In Frontiers in Environmental Science
(Vol. 2, Issue APR). Frontiers Media S.A. https://doi.org/10.3389/fenvs.2014.00007

17. Campbell, J. E., Lobell, D. B., Genova, R. C., & Field, C. B. (2008). The global potential of bioenergy on
abandoned agriculture lands. Environmental Science and Technology, 42(15), 5791–5794.
https://doi.org/10.1021/es800052w

18. Cohen-Shacham, E., Walters, G., Janzen, C., & Maginnis, S. (Eds.). (2016). Nature-based solutions to address
global societal challenges. IUCN International Union for Conservation of Nature.
https://doi.org/10.2305/IUCN.CH.2016.13.en

19. Cornelia Rumpel, Farshad Amiraslani, Lydie-Stella Koutika, Pete Smith, David Whitehead, & Eva
Wollenberg. (2018). Put more carbon in soils to meet Paris climate pledges. Nature, 564, 32–34.
https://doi.org/10.1038/d41586-018-07587-4

20. Correction for Sanderman et al., Soil carbon debt of 12,000 years of human land use. (2018). Proceedings of
the National Academy of Sciences, 115(7). https://doi.org/10.1073/pnas.1800925115

21. Cramer, V. A., Hobbs, R. J., & Standish, R. J. (2008). What’s new about old fields? Land abandonment and
ecosystem assembly. Trends in Ecology & Evolution, 23(2), 104–112.
https://doi.org/10.1016/J.TREE.2007.10.005

22. de Vries, W. (2018). Soil carbon 4 per mille: a good initiative but let’s manage not only the soil but also the
expectations: Comment on Minasny et al. (2017) Geoderma 292: 59–86. In Geoderma.
https://doi.org/10.1016/j.geoderma.2017.05.023

23. Deng, L., Liu, G. bin, & Shangguan, Z. ping. (2014a). Land-use conversion and changing soil carbon stocks
in China’s “Grain-for-Green” Program: A synthesis. Global Change Biology.
https://doi.org/10.1111/gcb.12508

24. Deng, L., Shangguan, Z. ping, & Sweeney, S. (2014b). “Grain for green” driven land use change and carbon
sequestration on the Loess Plateau, China. Scientific Reports, 4, 7039. https://doi.org/10.1038/srep07039

25. Deng, L., Zhu, G., Tang, Z., & Shangguan, Z. (2016). Global patterns of the effects of land-use changes on
soil carbon stocks. Global Ecology and Conservation, 5, 127–138.
https://doi.org/10.1016/J.GECCO.2015.12.004

26. Don, A., Schumacher, J., & Freibauer, A. (2011). Impact of tropical land-use change on soil organic carbon
stocks - a meta-analysis. In Global Change Biology. https://doi.org/10.1111/j.1365-2486.2010.02336.x

27. Dymov, A. A., Dubrovskiy, Y. A., & Startsev, V. v. (2018). Postagrogenic development of Retisols in the
middle taiga subzone of European Russia (Komi Republic). Land Degradation and Development, 29(3), 495–
505. https://doi.org/10.1002/ldr.2881

28. Dynarski, K. A., Bossio, D. A., & Scow, K. M. (2020). Dynamic Stability of Soil Carbon: Reassessing the
“Permanence” of Soil Carbon Sequestration. Frontiers in Environmental Science, 8.
https://doi.org/10.3389/fenvs.2020.514701

29. Eglin, T., Ciais, P., Piao, S. L., Barre, P., Bellassen, V., Cadule, P., Chenu, C., Gasser, T., Koven, C.,
Reichstein, M., & Smith, P. (2010). Historical and future perspectives of global soil carbon response to
climate and land-use changes. Tellus B: Chemical and Physical Meteorology, 62(5), 700–718.
https://doi.org/10.1111/j.1600-0889.2010.00499.x

30. Erb, K.-H., Luyssaert, S., Meyfroidt, P., Pongratz, J., Don, A., Kloster, S., Kuemmerle, T., Fetzel, T., Fuchs,
R., Herold, M., Haberl, H., Jones, C. D., Marín-Spiotta, E., McCallum, I., Robertson, E., Seufert, V., Fritz,
S., Valade, A., Wiltshire, A., & Dolman, A. J. (2017). Land management: data availability and process
understanding for global change studies. Global Change Biology, 23(2), 512–533.
https://doi.org/10.1111/gcb.13443

34
CHAPTER I: Introduction

31. FAO. (2006). The Role of Agriculture and Rural Development in Revitalizing Abandoned/Depopulated
Areas.

32. FAO, & ITPS. (2021). Recarbonizing Global Soils - A technical manual of recommended sustainable soil
management. Volume 3: Cropland, Grassland, Integrated systems, and farming approaches - Practices
Overview (Vol. 3). FAO. https://doi.org/10.4060/cb6595en

33. Fayet, C. M. J., Reilly, K. H., van Ham, C., & Verburg, P. H. (2022). What is the future of abandoned
agricultural lands? A systematic review of alternative trajectories in Europe. Land Use Policy, 112, 105833.
https://doi.org/10.1016/j.landusepol.2021.105833

34. Folberth, C., Skalský, R., Moltchanova, E., Balkovič, J., Azevedo, L. B., Obersteiner, M., & van der Velde,
M. (2016). Uncertainty in soil data can outweigh climate impact signals in global crop yield simulations.
Nature Communications, 7(1), 11872. https://doi.org/10.1038/ncomms11872

35. Foley, J. A., DeFries, R., Asner, G. P., Barford, C., Bonan, G., Carpenter, S. R., Chapin, F. S., Coe, M. T.,
Daily, G. C., Gibbs, H. K., Helkowski, J. H., Holloway, T., Howard, E. A., Kucharik, C. J., Monfreda, C.,
Patz, J. A., Prentice, I. C., Ramankutty, N., & Snyder, P. K. (2005). Global Consequences of Land Use.
Science, 309(5734), 570–574. https://doi.org/10.1126/science.1111772

36. Friedlingstein, P., Jones, M. W., O’Sullivan, M., Andrew, R. M., Bakker, D. C. E., Hauck, J., le Quéré, C.,
Peters, G. P., Peters, W., Pongratz, J., Sitch, S., Canadell, J. G., Ciais, P., Jackson, R. B., Alin, S. R., Anthoni,
P., Bates, N. R., Becker, M., Bellouin, N., … Zeng, J. (2022). Global Carbon Budget 2021. Earth System
Science Data, 14(4), 1917–2005. https://doi.org/10.5194/essd-14-1917-2022

37. Galloway, J. N., Townsend, A. R., Erisman, J. W., Bekunda, M., Cai, Z., Freney, J. R., Martinelli, L. A.,
Seitzinger, S. P., & Sutton, M. A. (2008). Transformation of the Nitrogen Cycle: Recent Trends, Questions,
and Potential Solutions. Science, 320(5878), 889–892. https://doi.org/10.1126/science.1136674

38. Gleixner, G., Poirier, N., Bol, R., & Balesdent, J. (2002). Molecular dynamics of organic matter in a cultivated
soil. Organic Geochemistry, 33(3), 357–366. https://doi.org/10.1016/S0146-6380(01)00166-8

39. Goldewijk, K. K., Beusen, A., van Drecht, G., & de Vos, M. (2011). The HYDE 3.1 spatially explicit database
of human-induced global land-use change over the past 12,000 years. Global Ecology and Biogeography,
20(1), 73–86. https://doi.org/10.1111/j.1466-8238.2010.00587.x

40. Griscom, B. W., Adams, J., Ellis, P. W., Houghton, R. A., Lomax, G., Miteva, D. A., Schlesinger, W. H.,
Shoch, D., Siikamäki, J. v., Smith, P., Woodbury, P., Zganjar, C., Blackman, A., Campari, J., Conant, R. T.,
Delgado, C., Elias, P., Gopalakrishna, T., Hamsik, M. R., … Fargione, J. (2017). Natural climate solutions.
Proceedings of the National Academy of Sciences, 114(44), 11645–11650.
https://doi.org/10.1073/pnas.1710465114

41. Guo, L. B., & Gifford, R. M. (2002). Soil carbon stocks and land use change: A meta analysis. Global Change
Biology. https://doi.org/10.1046/j.1354-1013.2002.00486.x

42. Gurevitch, J., Curtis, P. S., & Jones, M. H. (2001). Meta-analysis in ecology. Advances in Ecological
Research, 32, 199–247. https://doi.org/10.1016/S0065-2504(01)32013-5

43. Gutiérrez Rodríguez, L., Hogarth, N., Zhou, W., Putzel, L., Xie, C., & Zhang, K. (2015). Socioeconomic and
environmental effects of China’s Conversion of Cropland to Forest Program after 15 years: A systematic
review protocol. In Environmental Evidence. https://doi.org/10.1186/s13750-015-0033-8

44. Haberl, H., Erb, K. H., Krausmann, F., Gaube, V., Bondeau, A., Plutzar, C., Gingrich, S., Lucht, W., &
Fischer-Kowalski, M. (2007). Quantifying and mapping the human appropriation of net primary production
in earth’s terrestrial ecosystems. Proceedings of the National Academy of Sciences, 104(31), 12942–12947.
https://doi.org/10.1073/pnas.0704243104

45. Haberl, H., Erb, K.-H., & Krausmann, F. (2014). Human Appropriation of Net Primary Production: Patterns,
Trends, and Planetary Boundaries. Annual Review of Environment and Resources, 39(1), 363–391.
https://doi.org/10.1146/annurev-environ-121912-094620

35
CHAPTER I: Introduction

46. Hedges, L. v., Gurevitch, J., & Curtis, P. S. (1999). The meta-analysis of response ratios in experimental
ecology. Ecology.

47. Heidorn, P. B. (2008). Shedding Light on the Dark Data in the Long Tail of Science. Library Trends, 57(2),
280–299. https://doi.org/10.1353/lib.0.0036

48. Heinimann, A., Mertz, O., Frolking, S., Christensen, A. E., Hurni, K., Sedano, F., Chini, L. P., Sahajpal, R.,
Hansen, M., & Hurtt, G. (2017). A global view of shifting cultivation: Recent, current, and future extent.
PLoS ONE, 12(9). https://doi.org/10.1371/journal.pone.0184479

49. Hendriks, C. M. J., Stoorvogel, J. J., & Claessens, L. (2016). Exploring the challenges with soil data in
regional land use analysis. Agricultural Systems, 144, 9–21. https://doi.org/10.1016/j.agsy.2016.01.007

50. Henebry, G. M. (2009). Carbon in idle croplands. Nature, 457(7233), 1089–1090.


https://doi.org/10.1038/4571089a

51. Hoekstra, A. Y., & Wiedmann, T. O. (2014). Humanity’s unsustainable environmental footprint. Science,
344(6188), 1114–1117. https://doi.org/10.1126/science.1248365

52. Hong, S., Yin, G., Piao, S., Dybzinski, R., Cong, N., Li, X., Wang, K., Peñuelas, J., Zeng, H., & Chen, A.
(2020). Divergent responses of soil organic carbon to afforestation. Nature Sustainability, 3(9), 694–700.
https://doi.org/10.1038/s41893-020-0557-y

53. Houghton, R. A., House, J. I., Pongratz, J., van der Werf, G. R., DeFries, R. S., Hansen, M. C., le Quéré, C.,
& Ramankutty, N. (2012). Carbon emissions from land use and land-cover change. Biogeosciences, 9(12),
5125–5142. https://doi.org/10.5194/bg-9-5125-2012

54. Huggett, R. J. (1998). Soil chronosequences, soil development, and soil evolution: A critical review. In
Catena. https://doi.org/10.1016/S0341-8162(98)00053-8

55. IPCC. (2019). Climate Change and Land: an IPCC special report on climate change, desertification, land
degradation, sustainable land management, food security, and greenhouse gas fluxes in terrestrial ecosystems
(P. R. Shukla, J. Skea, E. Calvo Buendia, V. Masson-Delmotte, O. Pörtner, D. C. Roberts, P. Zhai, R. Slade,
S. Connors, R. van Diemem, M. Ferrat, E. Haughey, S. Luz, S. Neogi, M. Pathak, J. Petzold, J. Portugal
Pereira, E. Vyas, E. Huntley, … J. Mallley, Eds.). In press.

56. IPCC. (2021). Summary for Policymakers. In V. Masson-Delmotte, P. Zhai, A. Pirani, S. L. Connors, C.
Péan, S. Berger, N. Caud, Y. Chen, L. Goldfarb, M. I. Gomis, M. Huang, K. Leitzell, E. Lonnoy, J. B. R.
Matthews, T. K. Maycock, T. Waterfield, O. Yelekçi, R. Yu, & B. Zhou (Eds.), Climate Change 2021: The
Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the
Intergovernmental Panel on Climate Change. Cambridge University Press. In Press.

57. Isbell, F., Gonzalez, A., Loreau, M., Cowles, J., Díaz, S., Hector, A., Mace, G. M., Wardle, D. A., O’Connor,
M. I., Duffy, J. E., Turnbull, L. A., Thompson, P. L., & Larigauderie, A. (2017). Linking the influence and
dependence of people on biodiversity across scales. Nature, 546(7656), 65–72.
https://doi.org/10.1038/nature22899

58. Kalinina, O., Goryachkin, S. V., Lyuri, D. I., & Giani, L. (2015). Post-agrogenic development of vegetation,
soils, and carbon stocks under self-restoration in different climatic zones of European Russia. CATENA, 129,
18–29. https://doi.org/10.1016/J.CATENA.2015.02.016

59. Kämpf, I., Hölzel, N., Störrle, M., Broll, G., & Kiehl, K. (2016). Potential of temperate agricultural soils for
carbon sequestration: A meta-analysis of land-use effects. Science of The Total Environment, 566–567, 428–
435. https://doi.org/10.1016/J.SCITOTENV.2016.05.067

60. Katayama, N., Osawa, T., Amano, T., & Kusumoto, Y. (2015). Are both agricultural intensification and
farmland abandonment threats to biodiversity? A test with bird communities in paddy-dominated landscapes.
Agriculture, Ecosystems & Environment, 214, 21–30. https://doi.org/10.1016/J.AGEE.2015.08.014

36
CHAPTER I: Introduction

61. Keesstra, S., Nunes, J., Novara, A., Finger, D., Avelar, D., Kalantari, Z., & Cerdà, A. (2018). The superior
effect of nature based solutions in land management for enhancing ecosystem services. Science of The Total
Environment, 610–611, 997–1009. https://doi.org/10.1016/j.scitotenv.2017.08.077

62. Knops, J. M. H., & Tilman, D. (2000). Dynamics of soil nitrogen and carbon accumulation for 61 years after
agricultural abandonment. Ecology , 81(1), 88–98.

63. Koch, A., Brierley, C., Maslin, M. M., & Lewis, S. L. (2019). Earth system impacts of the European arrival
and Great Dying in the Americas after 1492. Quaternary Science Reviews, 207, 13–36.
https://doi.org/10.1016/J.QUASCIREV.2018.12.004

64. Kögel‐Knabner, I., Zech, W., & Hatcher, P. G. (1988). Chemical composition of the organic matter in forest
soils: The humus layer. Zeitschrift Für Pflanzenernährung Und Bodenkunde, 151(5), 331–340.
https://doi.org/10.1002/jpln.19881510512

65. Krause, A., Haverd, V., Poulter, B., Anthoni, P., Quesada, B., Rammig, A., & Arneth, A. (2019). Multimodel
Analysis of Future Land Use and Climate Change Impacts on Ecosystem Functioning. Earth’s Future, 7(7),
833–851. https://doi.org/10.1029/2018EF001123

66. Kuemmerle, T., Olofsson, P., Chaskovskyy, O., Baumann, M., Ostapowicz, K., Woodcock, C. E., Houghton,
R. A., Hostert, P., Keeton, W. S., & Radeloff, V. C. (2011). Post-Soviet farmland abandonment, forest
recovery, and carbon sequestration in western Ukraine. Global Change Biology.
https://doi.org/10.1111/j.1365-2486.2010.02333.x

67. Kuhn, A., & Heckelei, T. (2010). Anthroposphere. In Impacts of Global Change on the Hydrological Cycle
in West and Northwest Africa (pp. 282–341). Springer Berlin Heidelberg. https://doi.org/10.1007/978-3-642-
12957-5_8

68. Kurganova, I., Lopes de Gerenyu, V., & Kuzyakov, Y. (2015). Large-scale carbon sequestration in post-
agrogenic ecosystems in Russia and Kazakhstan. CATENA, 133, 461–466.
https://doi.org/10.1016/J.CATENA.2015.06.002

69. Kurganova, I., Lopes de Gerenyu, V., Six, J., & Kuzyakov, Y. (2014). Carbon cost of collective farming
collapse in Russia. Global Change Biology. https://doi.org/10.1111/gcb.12379

70. Laganière, J., Angers, D. A., & Paré, D. (2010). Carbon accumulation in agricultural soils after afforestation:
A meta-analysis. Global Change Biology, 16(1), 439–453. https://doi.org/10.1111/j.1365-2486.2009.01930.x

71. Lal, R. (2004a). Soil Carbon Sequestration Impacts on Global Climate Change and Food Security. Science,
304(5677), 1623–1627. https://doi.org/10.1126/science.1097396

72. Lal, R. (2004b). Soil carbon sequestration to mitigate climate change. Geoderma, 123(1–2), 1–22.
https://doi.org/10.1016/J.GEODERMA.2004.01.032

73. Lal, R. (2008). Carbon sequestration. In Philosophical Transactions of the Royal Society B: Biological
Sciences (Vol. 363, Issue 1492, pp. 815–830). Royal Society. https://doi.org/10.1098/rstb.2007.2185

74. Lal, R. (2013). Soil carbon management and climate change. Carbon Management, 4(4), 439–462.
https://doi.org/10.4155/cmt.13.31

75. Lal, R., Kimble, J., & Follett, R. F. (1998). Pedospheric Processes and the Carbon Cycle. In Soil Processes
and the Carbon Cycle (1st ed., pp. 1–8). CRC Press.

76. Lavallee, J. M., Soong, J. L., & Cotrufo, M. F. (2020). Conceptualizing soil organic matter into particulate
and mineral‐associated forms to address global change in the 21st century. Global Change Biology, 26(1),
261–273. https://doi.org/10.1111/gcb.14859

77. Lehmann, J., & Kleber, M. (2015). The contentious nature of soil organic matter. In Nature.
https://doi.org/10.1038/nature16069

37
CHAPTER I: Introduction

78. Leirpoll, M. E., Næss, J. S., Cavalett, O., Dorber, M., Hu, X., & Cherubini, F. (2021). Optimal combination
of bioenergy and solar photovoltaic for renewable energy production on abandoned cropland. Renewable
Energy, 168, 45–56. https://doi.org/10.1016/J.RENENE.2020.11.159

79. Lesiv, M., Schepaschenko, D., Moltchanova, E., Bun, R., Dürauer, M., Prishchepov, A. v., Schierhorn, F.,
Estel, S., Kuemmerle, T., Alcántara, C., Kussul, N., Shchepashchenko, M., Kutovaya, O., Martynenko, O.,
Karminov, V., Shvidenko, A., Havlik, P., Kraxner, F., See, L., & Fritz, S. (2018). Spatial distribution of arable
and abandoned land across former Soviet Union countries. Scientific Data.
https://doi.org/10.1038/sdata.2018.56

80. Lewis, S. L., & Maslin, M. A. (2015). Defining the Anthropocene. Nature, 519(7542), 171–180.
https://doi.org/10.1038/nature14258

81. Li, S., & Li, X. (2017). Global understanding of farmland abandonment: A review and prospects. Journal of
Geographical Sciences. https://doi.org/10.1007/s11442-017-1426-0

82. Li, W., Ciais, P., Guenet, B., Peng, S., Chang, J., Chaplot, V., Khudyaev, S., Peregon, A., Piao, S., Wang, Y.,
& Yue, C. (2018). Temporal response of soil organic carbon after grassland‐related land‐use change. Global
Change Biology. https://doi.org/10.1111/gcb.14328

83. Liu, X., Yang, T., Wang, Q., Huang, F., & Li, L. (2018). Dynamics of soil carbon and nitrogen stocks after
afforestation in arid and semi-arid regions: A meta-analysis. Science of The Total Environment, 618, 1658–
1664. https://doi.org/10.1016/J.SCITOTENV.2017.10.009

84. Lu, C., & Tian, H. (2017). Global nitrogen and phosphorus fertilizer use for agriculture production in the past
half century: shifted hot spots and nutrient imbalance. Earth System Science Data, 9(1), 181–192.
https://doi.org/10.5194/essd-9-181-2017

85. Lucas-Borja, M. E., Calsamiglia, A., Fortesa, J., García-Comendador, J., Lozano Guardiola, E., García-
Orenes, F., Gago, J., & Estrany, J. (2018). The role of wildfire on soil quality in abandoned terraces of three
Mediterranean micro-catchments. CATENA, 170, 246–256.
https://doi.org/10.1016/J.CATENA.2018.06.014

86. Marin-Spiotta, E., Silver, W. L., Swanston, C. W., & Ostertag, R. (2009). Soil organic matter dynamics during
80 years of reforestation of tropical pastures. Global Change Biology, 15(6), 1584–1597.
https://doi.org/10.1111/j.1365-2486.2008.01805.x

87. Mellor, P., Lord, R. A., João, E., Thomas, R., & Hursthouse, A. (2021). Identifying non-agricultural marginal
lands as a route to sustainable bioenergy provision - A review and holistic definition. Renewable and
Sustainable Energy Reviews, 135, 110220. https://doi.org/10.1016/j.rser.2020.110220

88. Minasny, B., Malone, B. P., McBratney, A. B., Angers, D. A., Arrouays, D., Chambers, A., Chaplot, V.,
Chen, Z. S., Cheng, K., Das, B. S., Field, D. J., Gimona, A., Hedley, C. B., Hong, S. Y., Mandal, B., Marchant,
B. P., Martin, M., McConkey, B. G., Mulder, V. L., … Winowiecki, L. (2017). Soil carbon 4 per mille. In
Geoderma. https://doi.org/10.1016/j.geoderma.2017.01.002

89. Minasny, B., McBratney, A. B., Wadoux, A. M. J. C., Akoeb, E. N., & Sabrina, T. (2020). Precocious 19th
century soil carbon science. Geoderma Regional, 22. https://doi.org/10.1016/j.geodrs.2020.e00306

90. Munson, S. M., Lauenroth, W. K., & Burke, I. C. (2012). Soil carbon and nitrogen recovery on semiarid
Conservation Reserve Program lands. Journal of Arid Environments, 79, 25–31.
https://doi.org/10.1016/j.jaridenv.2011.11.027

91. Navarro, L. M., & Pereira, H. M. (2012). Rewilding abandoned landscapes in Europe. In Rewilding European
Landscapes (pp. 3–23). Springer International Publishing. https://doi.org/10.1007/978-3-319-12039-3_1

92. Novick, K. A., Biederman, J. A., Desai, A. R., Litvak, M. E., Moore, D. J. P., Scott, R. L., & Torn, M. S.
(2018). The AmeriFlux network: A coalition of the willing. Agricultural and Forest Meteorology, 249, 444–
456. https://doi.org/10.1016/j.agrformet.2017.10.009

38
CHAPTER I: Introduction

93. Oberč, B. P., & Arroyo Schnell, A. (2020). Approaches to sustainable agriculture: exploring the pathways
towards the future of farming. IUCN, International Union for Conservation of Nature.
https://doi.org/10.2305/IUCN.CH.2020.07.en

94. Paustian, K., Lehmann, J., Ogle, S., Reay, D., Robertson, G. P., & Smith, P. (2016). Climate-smart soils. In
Nature (Vol. 532, Issue 7597, pp. 49–57). Nature Publishing Group. https://doi.org/10.1038/nature17174

95. Peñuelas, J., Poulter, B., Sardans, J., Ciais, P., van der Velde, M., Bopp, L., Boucher, O., Godderis, Y.,
Hinsinger, P., Llusia, J., Nardin, E., Vicca, S., Obersteiner, M., & Janssens, I. A. (2013). Human-induced
nitrogen–phosphorus imbalances alter natural and managed ecosystems across the globe. Nature
Communications, 4(1), 2934. https://doi.org/10.1038/ncomms3934

96. Poeplau, C., Don, A., Vesterdal, L., Leifeld, J., van Wesemael, B., Schumacher, J., & Gensior, A. (2011).
Temporal dynamics of soil organic carbon after land-use change in the temperate zone - carbon response
functions as a model approach. In Global Change Biology. https://doi.org/10.1111/j.1365-2486.2011.02408.x

97. Poore, J. A. C. (2016). Call for conservation: Abandoned pasture. Science, 351(6269), 132.

98. Post, W. M., & Kwon, K. C. (2000). Soil carbon sequestration and land‐use change: processes and potential.
Global Change Biology, 6(3). https://doi.org/10.1046/j.1365-2486.2000.00308.x

99. Potapov, P., Turubanova, S., Hansen, M. C., Tyukavina, A., Zalles, V., Khan, A., Song, X.-P., Pickens, A.,
Shen, Q., & Cortez, J. (2022). Global maps of cropland extent and change show accelerated cropland
expansion in the twenty-first century. Nature Food, 3(1), 19–28. https://doi.org/10.1038/s43016-021-00429-
z

100. Poulton, P., Johnston, J., Macdonald, A., White, R., & Powlson, D. (2018). Major limitations to achieving “4
per 1000” increases in soil organic carbon stock in temperate regions: Evidence from long-term experiments
at Rothamsted Research, United Kingdom. Global Change Biology. https://doi.org/10.1111/gcb.14066

101. Pribyl, D. W. (2010). A critical review of the conventional SOC to SOM conversion factor. Geoderma, 156(3–
4), 75–83. https://doi.org/10.1016/j.geoderma.2010.02.003

102. Prishchepov, A. v., Schierhorn, F., & Löw, F. (2021). Unraveling the Diversity of Trajectories and Drivers
of Global Agricultural Land Abandonment. Land, 10(2), 97. https://doi.org/10.3390/land10020097

103. Queiroz, C., Beilin, R., Folke, C., & Lindborg, R. (2014). Farmland abandonment: Threat or opportunity for
biodiversity conservation? A global review. In Frontiers in Ecology and the Environment.
https://doi.org/10.1890/120348

104. Quesada, B., Arneth, A., Robertson, E., & de Noblet-Ducoudré, N. (2018). Potential strong contribution of
future anthropogenic land-use and land-cover change to the terrestrial carbon cycle. Environmental Research
Letters, 13(6), 064023. https://doi.org/10.1088/1748-9326/aac4c3

105. Ramankutty, N., Evan, A. T., Monfreda, C., & Foley, J. A. (2008). Farming the planet: 1. Geographic
distribution of global agricultural lands in the year 2000. Global Biogeochemical Cycles, 22(1), n/a-n/a.
https://doi.org/10.1029/2007GB002952

106. Ramankutty, N., Mehrabi, Z., Waha, K., Jarvis, L., Kremen, C., Herrero, M., & Rieseberg, L. H. (2018).
Trends in Global Agricultural Land Use: Implications for Environmental Health and Food Security. Annual
Review of Plant Biology, 69(1), 789–815. https://doi.org/10.1146/annurev-arplant-042817-040256

107. Rasmussen, C., Heckman, K., Wieder, W. R., Keiluweit, M., Lawrence, C. R., Berhe, A. A., Blankinship, J.
C., Crow, S. E., Druhan, J. L., Hicks Pries, C. E., Marin-Spiotta, E., Plante, A. F., Schädel, C., Schimel, J. P.,
Sierra, C. A., Thompson, A., & Wagai, R. (2018). Beyond clay: towards an improved set of variables for
predicting soil organic matter content. Biogeochemistry, 137(3), 297–306. https://doi.org/10.1007/s10533-
018-0424-3

108. Rasse, D. P., Dignac, M.-F., Bahri, H., Rumpel, C., Mariotti, A., & Chenu, C. (2006). Lignin turnover in an
agricultural field: from plant residues to soil-protected fractions. European Journal of Soil Science, 57(4),
530–538. https://doi.org/10.1111/j.1365-2389.2006.00806.x

39
CHAPTER I: Introduction

109. Rey Benayas, J. M., Martins, A., Nicolau, J. M., & Schulz, J. J. (2007). Abandonment of agricultural land:
an overview of drivers and consequences. CAB Reviews: Perspectives in Agriculture, Veterinary Science,
Nutrition and Natural Resources. https://doi.org/10.1079/PAVSNNR20072057

110. Ritchie, H. (2022, May 30). After millennia of agricultural expansion, the world has passed ‘peak agricultural
land.’ Our World in Data. https://ourworldindata.org/peak-agriculture-land

111. Robles, M. D., & Burke, I. C. (1998). Soil Organic Matter Recovery on Conservation Reserve Program Fields
in Southeastern Wyoming. Soil Science Society of America Journal, 62(3), 725–730.
https://doi.org/10.2136/sssaj1998.03615995006200030026x

112. Rodrigues, L., Fohrafellner, J., Hardy, B., Huyghebaert, B., & Leifeld, J., (2021). Towards climate-smart
sustainable management of agricultural soils. Deliverable 2.3 Synthesis on estimates of achievable soil carbon
sequestration on agricutural land across Europe.

113. Ryo, M., Aguilar-Trigueros, C. A., Pinek, L., Muller, L. A. H., & Rillig, M. C. (2019). Basic Principles of
Temporal Dynamics. Trends in Ecology & Evolution, 34(8), 723–733.
https://doi.org/10.1016/j.tree.2019.03.007

114. Sanderman, J., Hengl, T., & Fiske, G. J. (2017). Soil carbon debt of 12,000 years of human land use.
Proceedings of the National Academy of Sciences. https://doi.org/10.1073/pnas.1706103114

115. Sarkar, D., Bungbungcha Meitei, Ch., Baishya, L. K., Das, A., Ghosh, S., Chongloi, K. L., & Rajkhowa, D.
(2015). Potential of fallow chronosequence in shifting cultivation to conserve soil organic carbon in northeast
India. CATENA, 135, 321–327. https://doi.org/10.1016/J.CATENA.2015.08.012

116. Schierhorn, F., Kastner, T., Kuemmerle, T., Meyfroidt, P., Kurganova, I., Prishchepov, A. v, Erb, K.-H.,
Houghton, R. A., & Müller, D. (2019). Large greenhouse gas savings due to changes in the post-Soviet food
systems. Environmental Research Letters, 14(6), 065009. https://doi.org/10.1088/1748-9326/ab1cf1

117. Schierhorn, F., Müller, D., Beringer, T., Prishchepov, A. v., Kuemmerle, T., & Balmann, A. (2013). Post-
Soviet cropland abandonment and carbon sequestration in European Russia, Ukraine, and Belarus. Global
Biogeochemical Cycles, 27(4), 1175–1185. https://doi.org/10.1002/2013GB004654

118. Schmidt, M. W. I., Torn, M. S., Abiven, S., Dittmar, T., Guggenberger, G., Janssens, I. A., Kleber, M., Kögel-
Knabner, I., Lehmann, J., Manning, D. A. C., Nannipieri, P., Rasse, D. P., Weiner, S., & Trumbore, S. E.
(2011). Persistence of soil organic matter as an ecosystem property. In Nature.
https://doi.org/10.1038/nature10386

119. Seddon, N., Turner, B., Berry, P., Chausson, A., & Girardin, C. (2019). Grounding nature-based climate
solutions in sound biodiversity science. Nature Climate Change, 9, 82–87. https://doi.org/10.1038/s41558-
019-0402-3

120. Shi, S., & Han, P. (2014). Estimating the soil carbon sequestration potential of China’s Grain for Green
Project. Global Biogeochemical Cycles, 28(11), 1279–1294. https://doi.org/10.1002/2014GB004924

121. Smith, P., Adams, J., Beerling, D. J., Beringer, T., Calvin, K. v., Fuss, S., Griscom, B., Hagemann, N.,
Kammann, C., Kraxner, F., Minx, J. C., Popp, A., Renforth, P., Vicente Vicente, J. L., & Keesstra, S. (2019).
Land-Management Options for Greenhouse Gas Removal and Their Impacts on Ecosystem Services and the
Sustainable Development Goals. Annual Review of Environment and Resources, 44(1), 255–286.
https://doi.org/10.1146/annurev-environ-101718-033129

122. Smith, P., Cotrufo, M. F., Rumpel, C., Paustian, K., Kuikman, P. J., Elliott, J. A., McDowell, R., Griffiths,
R. I., Asakawa, S., Bustamante, M., House, J. I., Sobocká, J., Harper, R., Pan, G., West, P. C., Gerber, J. S.,
Clark, J. M., Adhya, T., Scholes, R. J., & Scholes, M. C. (2015). Biogeochemical cycles and biodiversity as
key drivers of ecosystem services provided by soils. SOIL, 1(2), 665–685. https://doi.org/10.5194/soil-1-665-
2015

123. Song, X.-P., Hansen, M. C., Stehman, S. v., Potapov, P. v., Tyukavina, A., Vermote, E. F., & Townshend, J.
R. (2018). Global land change from 1982 to 2016. Nature, 560(7720), 639–643.
https://doi.org/10.1038/s41586-018-0411-9

40
CHAPTER I: Introduction

124. Subedi, Y. R., Kristiansen, P., & Cacho, O. (2022). Drivers and consequences of agricultural land
abandonment and its reutilisation pathways: A systematic review. Environmental Development, 42, 100681.
https://doi.org/10.1016/j.envdev.2021.100681

125. Sulman, B. N., Moore, J. A. M., Abramoff, R., Averill, C., Kivlin, S., Georgiou, K., Sridhar, B., Hartman,
M. D., Wang, G., Wieder, W. R., Bradford, M. A., Luo, Y., Mayes, M. A., Morrison, E., Riley, W. J., Salazar,
A., Schimel, J. P., Tang, J., & Classen, A. T. (2018). Multiple models and experiments underscore large
uncertainty in soil carbon dynamics. Biogeochemistry, 141(2), 109–123. https://doi.org/10.1007/s10533-018-
0509-z

126. Tharammal, T., Bala, G., Devaraju, N., & Nemani, R. (2019). A review of the major drivers of the terrestrial
carbon uptake: model-based assessments, consensus, and uncertainties. Environmental Research Letters,
14(9), 093005. https://doi.org/10.1088/1748-9326/ab3012

127. Tisdall, J. M., & Oades, J. M. (1982). Organic matter and water-stable aggregates in soils. Journal of Soil
Science, 33(2), 141–163. https://doi.org/10.1111/j.1365-2389.1982.tb01755.x

128. Torn, M. S., Trumbore, S. E., Chadwick, O. A., Vitousek, P. M., & Hendricks, D. M. (1997). Mineral control
of soil organic carbon storage and turnover. Nature, 389(6647), 170–173. https://doi.org/10.1038/38260

129. Trivedi, P., Singh, B. P., & Singh, B. K. (2018). Soil Carbon: Introduction, Importance, Status, Threat, and
Mitigation. In Soil Carbon Storage (pp. 1–28). Elsevier. https://doi.org/10.1016/B978-0-12-812766-7.00001-
9

130. Turner, B. L., Lambin, E. F., & Reenberg, A. (2007). The emergence of land change science for global
environmental change and sustainability. Proceedings of the National Academy of Sciences, 104(52), 20666–
20671. https://doi.org/10.1073/pnas.0704119104

131. Ustaoglu, E., & Collier, M. J. (2018). Farmland abandonment in Europe: an overview of drivers,
consequences, and assessment of the sustainability implications. Environmental Reviews, 26(4), 396–416.
https://doi.org/10.1139/er-2018-0001

132. Vermeulen, S., Bossio, D., Lehmann, J., Luu, P., Paustian, K., Webb, C., Augé, F., Bacudo, I., Baedeker, T.,
Havemann, T., Jones, C., King, R., Reddy, M., Sunga, I., von Unger, M., & Warnken, M. (2019). A global
agenda for collective action on soil carbon. In Nature Sustainability (Vol. 2, Issue 1, pp. 2–4). Nature
Publishing Group. https://doi.org/10.1038/s41893-018-0212-z

133. Vilà-Cabrera, A., Espelta, J. M., Vayreda, J., & Pino, J. (2017). “New Forests” from the Twentieth Century
are a Relevant Contribution for C Storage in the Iberian Peninsula. Ecosystems.
https://doi.org/10.1007/s10021-016-0019-6

134. Vines, T. H., Albert, A. Y. K., Andrew, R. L., Débarre, F., Bock, D. G., Franklin, M. T., Gilbert, K. J., Moore,
J.-S., Renaut, S., & Rennison, D. J. (2014). The Availability of Research Data Declines Rapidly with Article
Age. Current Biology, 24(1), 94–97. https://doi.org/10.1016/j.cub.2013.11.014

135. Vuichard, N., Ciais, P., Belelli, L., Smith, P., & Valentini, R. (2008). Carbon sequestration due to the
abandonment of agriculture in the former USSR since 1990. Global Biogeochemical Cycles.
https://doi.org/10.1029/2008GB003212

136. Walker, L. R., Wardle, D. A., Bardgett, R. D., & Clarkson, B. D. (2010). The use of chronosequences in
studies of ecological succession and soil development. Journal of Ecology. https://doi.org/10.1111/j.1365-
2745.2010.01664.x

137. Wertebach, T. M., Hölzel, N., Kämpf, I., Yurtaev, A., Tupitsin, S., Kiehl, K., Kamp, J., & Kleinebecker, T.
(2017). Soil carbon sequestration due to post-Soviet cropland abandonment: estimates from a large-scale soil
organic carbon field inventory. Global Change Biology, 23(9), 3729–3741.
https://doi.org/10.1111/gcb.13650

138. White, R. E., Davidson, B., Lam, S. K., & Chen, D. (2018). A critique of the paper ‘Soil carbon 4 per mille’’
by Minasny et al. (2017).’ In Geoderma. https://doi.org/10.1016/j.geoderma.2017.05.025

41
CHAPTER I: Introduction

139. Wieder, W. R., Hartman, M. D., Sulman, B. N., Wang, Y., Koven, C. D., & Bonan, G. B. (2018). Carbon
cycle confidence and uncertainty: Exploring variation among soil biogeochemical models. Global Change
Biology, 24(4), 1563–1579. https://doi.org/10.1111/gcb.13979

140. Winkler, K., Fuchs, R., Rounsevell, M., & Herold, M. (2021). Global land use changes are four times greater
than previously estimated. Nature Communications, 12(1), 2501. https://doi.org/10.1038/s41467-021-22702-
2

141. Yang, Y., Hobbie, S. E., Hernandez, R. R., Fargione, J., Grodsky, S. M., Tilman, D., Zhu, Y.-G., Luo, Y.,
Smith, T. M., Jungers, J. M., Yang, M., & Chen, W.-Q. (2020). Restoring Abandoned Farmland to Mitigate
Climate Change on a Full Earth. One Earth, 3(2), 176–186. https://doi.org/10.1016/j.oneear.2020.07.019

142. Yin, H., Brandão, A., Buchner, J., Helmers, D., Iuliano, B. G., Kimambo, N. E., Lewińska, K. E., Razenkova,
E., Rizayeva, A., Rogova, N., Spawn, S. A., Xie, Y., & Radeloff, V. C. (2020). Monitoring cropland
abandonment with Landsat time series. Remote Sensing of Environment, 246, 111873.
https://doi.org/10.1016/j.rse.2020.111873

143. Yin, H., Butsic, V., Buchner, J., Kuemmerle, T., Prishchepov, A. v., Baumann, M., Bragina, E. v., Sayadyan,
H., & Radeloff, V. C. (2019). Agricultural abandonment and re-cultivation during and after the Chechen Wars
in the northern Caucasus. Global Environmental Change, 55, 149–159.
https://doi.org/10.1016/j.gloenvcha.2019.01.005

144. Yin, H., Prishchepov, A. v., Kuemmerle, T., Bleyhl, B., Buchner, J., & Radeloff, V. C. (2018). Mapping
agricultural land abandonment from spatial and temporal segmentation of Landsat time series. Remote
Sensing of Environment, 210, 12–24. https://doi.org/10.1016/j.rse.2018.02.050

145. Zomer, R. J., Bossio, D. A., Sommer, R., & Verchot, L. v. (2017). Global Sequestration Potential of Increased
Organic Carbon in Cropland Soils. Scientific Reports. https://doi.org/10.1038/s41598-017-15794-8

42
CHAPTER I: Introduction

1.8 Supplementary materials


Table 2. Typically used delineation of SOM fractions and SOC pools according to the previous
SOM conceptual paradigm.
SOM fractions Residence Time Characteristics
Dissolved Minutes to days • < 45µm
• Root exudates, simple sugars, by-products of
decomposition
• < 5% of SOM
Particulate 2–50 years • 53µm–2mm
• Fresh and decomposing plant and animal matter
• 2–25% of SOM
Humus 10s–100s years • < 53µm
• Longer-lasting organic compounds under slow
decomposition
• > 50% of SOM
Resistant 100s–1000s years • < 53µm–2mm
• Comparatively inert, chemically resistant remnant
organic materials
• < 10% of SOM
SOC pools Residence Time Characteristics
Fast (labile or active) 1–2 years • Highly unstable and subject to decomposition
Intermediate 10–100 years • Microbially processed and partially stabilized on
mineral surfaces and protected within soil
aggregates
Slow (recalcitrant or 100– > 1000 • Highly stabilized
stable) years • Includes pyrogenic carbon

43
CHAPTER II: Management opportunities for soil
carbon sequestration following agricultural land
abandonment
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

2.1 Overview
The widespread historical and ongoing abandonment of agricultural lands worldwide presents
important opportunities for promoting climate change mitigation through carbon sequestration.
The default management outcome of abandonment is natural regeneration through ecological
succession. However, several different management strategies and new land uses for
abandoned agricultural lands have been recommended by the scientific community in recent
years. This paper reviews the foremost proposed strategies and compares their soil carbon
sequestration potentials. Six major categories have been proposed globally. Each proposal has
positive and negative outcomes depending on site-specific factors and management objectives.
Accordingly, no single strategy is ideal in all scenarios and a combination of strategies
addresses multiple rural development goals concurrently. A combination of passive and active
management techniques is the most effective approach for maximizing soil carbon
sequestration over large geographic scales, while other strategies can be designed to also
promote low-carbon land use practices and fossil fuel substitution. The implications of each
proposal highlighted here demonstrates the positive role that abandoned agricultural lands can
serve in climate change mitigation efforts, supporting policymakers tasked with planning the
future of regions undergoing abandonment.

46
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

2.2 Introduction
Vast expanses of previously cultivated lands around the world are currently undergoing the
processes of natural regeneration as a result of agricultural land abandonment (ALA) (Cramer
et al., 2008; Queiroz et al., 2014; Rey Benayas et al., 2007). Upwards of 472 Mha, or over half
of the land area of the USA, is estimated to have been abandoned over the last three centuries
globally (Campbell et al., 2008). With a further 280 Mha of shifting agriculture currently
undergoing cyclical abandonment (Heinimann et al., 2017), ALA as a land use change (LUC)
presents both important management challenges and opportunities.

While scientists and society debate whether or not to intervene with the regeneration process
post-abandonment to control the negative impacts of revegetation and promote desired
outcomes (Lasanta et al., 2015; Rey Benayas et al., 2008), alternative options are receiving
increasing attention (Abolina and Luzadis, 2015; Campbell et al., 2013; Hall, 2018; Knoke et
al., 2014; Pace Ricci and Conrad, 2018; Schröder et al., 2018; Smaliychuk et al., 2016). This
growing body of research has gone beyond describing ALA solely from a LUC perspective,
and has instead proposed management strategies to guide sustainable transitions into new land
use classifications. The recognized importance of land management for climate change
mitigation efforts (IPCC, 2019), alongside increasing demand for land resources, provides
ample reason to explore what these proposals are, the development goals they can meet, and
their implications for climate change.

The default management strategy post-abandonment is simply natural regeneration through


ecological succession, usually triggering both aboveground (i.e., plant biomass) and
belowground (i.e., soil) carbon accumulation (Silver et al., 2001). Considering that
agroecosystems are often depleted of soil organic carbon (SOC) (Lal et al., 2015), quantifying
the soil carbon sequestration potentials of each of these proposals is especially important for
assessing their climate change mitigation relevance. With recent advances in our understanding
of the consequences of ALA over the last two decades, it is also an opportune moment to review
how managing abandoned lands can contribute towards replenishing the soil carbon pool.

In light of these issues, this paper identifies and compares the most commonly proposed post-
abandonment management strategies globally through a review of the ALA literature. First, the
ecological and rural development features of each proposal are discussed in support of
policymakers tasked with planning the future of regions undergoing ALA. Second, the
proposals with soil carbon sequestration rates reported on abandoned or converted agricultural
lands synthesized over large geographic scales are compared to determine their relative

47
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

potential contributions towards climate change mitigation. With the reduced importance of
local environmental factors and differing agricultural histories at larger scales, active
management through restoration is hypothesized to produce the highest rates of SOC
accumulation post-LUC.

2.3 Methods
The following terms in various combinations were searched using ISI Web of Science, Scopus,
and Google Scholar: “agricultural land abandonment”, “farmland abandonment”, “cropland
abandonment”, “ex-arable lands”, “old fields”, “land abandonment”, “marginal land”,
“degraded land”, “land use change”, and “land management”. Search results were limited to
English language studies published over the last two decades. Each potential study was
assessed first by title to determine suitability, then by abstract to determine potential for land
management proposals, and lastly by main text to extract the most relevant information.
Reference lists were also reviewed and relevant grey literature was consulted when applicable.
As this is a global analysis, studies from all countries were considered. Management strategies
and new land uses for abandoned agricultural lands that were proposed by several authors
globally with case studies in at least two continents were selected for analysis. A second
literature search was performed to collect SOC data following implementation of each of the
selected management strategies. SOC sequestration rates were synthesized based on study area,
resulting in three classifications of large geographic scales featured in this review: country,
climatic zone, and global. All studies that were found reporting rates at the climatic zone or
global scale of any of the proposed management strategies on abandoned or directly converted
agricultural lands were included. Studies reporting data at the country scale were only included
if they featured comparably large sampling sizes with significant geographic extent. Rates
applying to the first three decades post-LUC and taken from topsoils (0–30 cm) where
preferentially selected.

2.4 Results & Discussion


2.4.1 Proposed management strategies

Although most of the ALA literature focuses primarily on drivers (Baumann et al., 2011; Díaz
et al., 2011; Osawa et al., 2016; Zhang et al., 2014), impacts (Cramer et al., 2008; Hanaček and
Rodríguez-Labajos, 2018; Lasanta et al., 2015), and spatial distribution (Estel et al., 2015;
Schulp et al., 2018; Verburg and Overmars, 2009; Wang et al., 2018; Yin et al., 2018),
researchers have also discussed potential management strategies and new land uses for

48
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

abandoned lands that take into account the unique social, environmental, and economic
conditions and needs of the stakeholders involved. Our literature review revealed six major
categories proposed by the scientific community globally (Figure 9). Three higher-tier
categories were identified based on the primary desired function of the new land use in which
all proposals could be further grouped (i.e., Naturalness, Multifunctionality, and Productivity).

Figure 9. Decision tree highlighting the major categories of proposed management strategies
and potential new land uses for abandoned agricultural lands globally. See Table 3 for a
summary.
At the farm and landscape scale, abandoned land is often perceived negatively and devalued,
motivating landowners and decision-makers to consider new management options (Abolina
and Luzadis, 2015; Benjamin et al., 2007; Ruskule et al., 2013; van der Zanden et al., 2018).
At the regional scale, a patchwork of several different adaptive management strategies likely
offers the best compromise for balancing ecological, social, and economic interests through
time (Merckx and Pereira, 2015; Pelorosso et al., 2011; Ruskule et al., 2013). Researchers

49
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

acknowledge that implementing the most suitable post-abandonment strategy will be based on
numerous variables, requiring increased efforts to address information gaps in the decision-
making process (Meli et al., 2017). To that end, the ecological and rural development features
of the six proposals are discussed below and summarized in Table 3.

2.4.1.1 Naturalness
Promoting naturalness on abandoned lands can be achieved with intensive human intervention
(i.e., active management) or without (i.e., passive management). Assisted afforestation and
grassland restoration via seeding and planting are commonly proposed active management
practices. On the other end of the spectrum, passive management practices aim to restore
naturalness through ecological succession (e.g., rewilding) (Perino et al., 2019).

Various factors play a significant role in the efficacy of both strategies, resulting in occasionally
differing impacts. The agricultural history, ecosystem characteristics, and local lithology of
abandoned lands may influence the rate at which vegetation recolonizes, biodiversity returns
to native states, and soil properties improve (Robledano-Aymerich et al., 2014). For example,
passive management on abandoned pastures in the Atlantic Forest region of Brazil resulted in
low aboveground biomass and low biodiversity, suggesting the need for enrichment plantings,
nucleation techniques (e.g., plantation in islands), and other forms of active management
(Sansevero et al., 2017). In the Amazon on the other hand, (Rocha et al., 2016) specifically
proposed natural regeneration as the ideal restoration strategy for abandoned pastures due to
the speed and intensity of revegetation observed.

Climate is indeed a key factor on whether active or passive techniques deliver intended
management outcomes. Throughout the tropics, rewilding has been found to outperform active
management for restoring biodiversity and vegetation structure, challenging the view that
natural regeneration has inferior conservation value in these ecosystems (Crouzeilles et al.,
2017). In semi-arid climates, afforestation of abandoned lands produced semi-natural soil
conditions after 40 years and formed important natural resource islands in the landscape, while
natural succession resulted in minor non-linear improvements leaving the soil prone to erosion
(Zethof et al., 2019). However, rewilding on abandoned lands is generally sufficient for faunal,
floral, and biogeochemical recovery across most climates globally (Meli et al., 2017).

The local differences in efficacy between rewilding and active restoration are not always easy
to discern. The variability of impacts of both strategies implies that decision-makers sometimes
need to consider different solutions for seemingly similar situations. For instance, slower

50
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

tropical forest regrowth rates on abandoned lands observed in Uganda, in comparison to South
America, has led researchers to propose silviculture plantations to create new sources of income
while allowing regeneration processes to occur under tree shelter (Chapman and Chapman,
1999). In the species-rich grasslands of Czech Republic, where ALA has been a significant
regional LUC over the last few decades, passive management for grassland restoration is
appropriate if the desired species are already present in ancient grasslands close by (Sojneková
and Chytrý, 2015). In situations where this might not be the case, specific mixtures of regional
seeds and hay transfer are recommended (Albert et al., 2019; Prach et al., 2014).

One solution often suggested is to employ both passive and active management within the same
region, usually involving localized restoration techniques rather than widespread rewilding
initiatives (Robledano-Aymerich et al., 2014). This would allow for the best of both worlds
and is ideal for climates where natural regeneration of forests and grasslands post-abandonment
would benefit from some level of human intervention. In the Mediterranean, for example,
tailor-made strategies at the farm or landscape scale (e.g., focusing planting efforts locally in
small groups of native trees to jumpstart natural regrowth) incorporated into regional mosaics
of differing landscapes may support biodiversity and ecosystem goals more efficiently
(Plieninger et al., 2014; Rey Benayas et al., 2008). From a productivity standpoint, rural
development subsidies for both rewilding on marginal lands and active biodiversity protection
with sustainable intensive agriculture on more productive lands is a practical approach (Merckx
and Pereira, 2015). Rewilding has been especially recommended in areas where the
socioeconomic activities of the landscape are no longer sustainable (Cerqueira et al., 2015).

2.4.1.2 Multifunctionality
Some of the proposals address a wider range of ecological and rural development goals,
reducing the need to combine strategies within the same landscape. The most often proposed
management strategies in the ALA literature in this category are forms of low-intensity
agriculture, agritourism and hobby farming. Low-intensity agriculture provides sustainable
food production and several other benefits valued in regions with marginal and abandoned
lands (Mander et al., 1999). In Europe, for instance, silvopastoral systems are proposed both
for their potential to provide beef while sequestering carbon (Hall, 2018) and their suitability
as low-intervention tools for limiting potential negative impacts of unmanaged revegetation
(e.g., wildfire risk) (Lasanta et al., 2015). In Central and South America, although government
policies have promoted natural regeneration to increase biodiversity, numerous studies instead
propose returning to, or conserve existing, low-intensity agricultural practices such as cacao

51
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

and coffee agroforestry and the traditional milpa system to achieve additional benefits (De
Beenhouwer et al., 2013; García-Frapolli et al., 2007; Jezeer et al., 2017; Londoño et al., 2017;
Queiroz et al., 2014; Santos et al., 2019; Tscharntke et al., 2011).

Further proposals to reclaim or maintain existing traditional low-intensity agroecosystems in


regions experiencing ALA have been made across diverse cultural and environmental settings
globally, such as the dehesa and montado of the Iberian Peninsula (Arosa et al., 2017; Garrido
et al., 2017) and the satoyama of Japan (Morimoto, 2011). Reclaiming traditional agricultural
landscapes such as these can also support local agritourism due to the improved landscape
aesthetics (i.e., tourist preferences for active over abandoned lands) (Zagaria et al., 2018), as
has been proposed for the Noto Peninsula in Japan (Chen et al., 2016) and Mediterranean
mountainous regions (Sayadi et al., 2009). Agritourism can be sustainable in regions
experiencing ALA while building socioeconomic resilience to climate induced threats to rural
livelihoods (Sanagustín-Fons et al., 2011; Valdivia and Barbieri, 2014).

Another approach with sociocultural benefits for rural development involves promoting hobby
and community social farming (Knapik, 2018), especially on abandoned lands located near
urban areas with large populations of interested citizens and engaged landowners (Pace Ricci
and Conrad, 2018). Further away from urban areas, hobby farming can also overcome
hereditary barriers associated with abandoned lands by bringing in new families to farm the
land and revive rural community life (Varotto and Lodatti, 2014).

2.4.1.3 Productivity
The third higher-tier category of proposals, which includes sustainable intensive agriculture
and bioenergy and renewables, involves generating economic value on abandoned lands
through producing goods (i.e., food and energy) sustainably. Calls for sustainable agricultural
intensification have been made worldwide to address increasing land pressures and demand for
food. Complementary to efforts to transition conventional agricultural lands, abandoned and
marginal agricultural lands are now receiving attention in this regard (Schröder et al., 2018).
Reclamation of abandoned lands in Ecuador, for example, has been proposed as a way to
increase food production, improve land allocation, decrease food prices, and prevent forest
conversion to agriculture (Knoke et al., 2013). For global dryland regions, reestablishment of
agriculture on abandoned or marginal lands inherently entails incorporating best management
practices from integrated soil fertility management and conservation/climate smart agriculture
(Shahid and Al-Shankiti, 2013). The greatest potential for sustainable intensification possibly

52
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

lies in post-Soviet states where significant recent reclamation has already occurred (Schierhorn
et al., 2013).

The basis of proposing bioenergy crops and renewables on abandoned lands is the argument
that competition for land with food production, one of the central constraints of large-scale
implementation, would no longer be a factor. These two energy production methods are already
promoted as ways to reduce greenhouse gas emissions and reliance on fossil fuels while
creating new development opportunities in the agricultural sector. Consequently, researchers
argue that their integration on marginal and abandoned lands greatly increases their
attractiveness and large-scale viability (Campbell et al., 2008; Edrisi and Abhilash, 2016;
Fargione et al., 2008; Milbrandt et al., 2014; Nijsen et al., 2012; Valentine et al., 2012;
Vuichard et al., 2009). Campbell et al., (2008) estimated the global potential of bioenergy on
abandoned lands, and found that it could easily satisfy the national energy demand of some
African nations, while reaching less than 10% of the demand for most countries in North
America, Europe, and Asia. Subsequent studies have argued that this option can still meet a
meaningful proportion of the primary energy demand of several countries across these
continents (Campbell et al., 2013; Liu et al., 2017; Nijsen et al., 2012).

Renewable energy technologies established on marginal and abandoned lands are also worthy
of consideration. In the USA, such installations could significantly increase domestic energy
supply even if only a fraction of the potential is realized, with solar power offering the highest
return (Milbrandt et al., 2014). Importantly, large-scale implementation of bio- or renewable
energy production on marginal and/or degraded lands is considered by some to be too costly
due to accessibility and soil productivity issues, implying that abandoned lands may be a more
suitable option (Baxter and Calvert, 2017; Bryngelsson and Lindgren, 2013; Shortall, 2013;
Swinton et al., 2011).

53
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

Table 3. Summary of proposed management strategies and potential new land uses for abandoned agricultural lands globally.

Ideal landscape
Proposal Purpose Reasons for proposal Relevant studies
conditions
Naturalness
• Increased aboveground • Degraded
Ecosystem carbon sequestration • High
(Knoke et al., 2014; Prach et al., 2014;
Active services • Reduced soil erosion ecological
Rey Benayas et al., 2008; Tomaz et al.,
Management: through • Potential return of large value
2013; Tremblay and Ouimet, 2013;
Restoration managed fauna • Low potential
Zethof et al., 2019)
restoration • Strategic management of for natural
forest regrowth regeneration
• Increased belowground • Degraded
Ecosystem carbon sequestration • High
(Basualdo et al., 2019; Crouzeilles et al.,
Passive services • Recovery of soil functions ecological
2017; Jiao et al., 2011; Meli et al., 2017;
Management: naturally • Potential return of large value
Navarro and Pereira, 2012; Regos et al.,
Rewilding without fauna • High potential
2016; Scott and Morgan, 2012)
intervention • Increased biodiversity for natural
• Low-cost regeneration
Multifunctionality
• Increased habitat
connectivity, biodiversity,
above- and belowground
(Arosa et al., 2017; Hall, 2018;
Sustainable carbon sequestration (vs.
• High Hombegowda et al., 2016; Jezeer et al.,
food conventional agriculture)
Low-intensity ecological 2017; Lasanta et al., 2015; Londoño et
production and • Recovery of soil functions
Agriculture value al., 2017; Morimoto, 2011; Santos et al.,
ecosystem • Reduced wildfire risk (with
• Arable soils 2019; Smith et al., 2013; Torralba et al.,
services grazing)
2016)
• Sustainable food production
• Rural development
opportunities

54
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

Ideal landscape
Proposal Purpose Reasons for proposal Relevant studies
conditions
• Association
• Social/ecological resilience
with traditional
to climate threats
Traditional/sm agricultural (Chen et al., 2016; Pace Ricci and
• Rejuvenate lands under
Agritourism & all-scale food practices Conrad, 2018; Sanagustín-Fons et al.,
inactive ownership
Hobby production and • Low 2011; Sayadi et al., 2009; Valdivia and
• Preserve cultural and
Farming rural ecological Barbieri, 2014; Varotto and Lodatti,
aesthetic landscapes
development value 2014; Zagaria et al., 2018)
• Rural development
• Near large
opportunities
populations
Productivity
• Alleviate land pressure/avoid
Sustainable deforestation (through use of • Low (Benjamin et al., 2007; Griffiths et al.,
Sustainable
large-scale abandoned lands) ecological 2013; Knoke et al., 2013; Schröder et al.,
Intensive
food • Recovery of some soil value 2018; Shahid and Al-Shankiti, 2013;
Agriculture
production functions • Arable soils Smaliychuk et al., 2016)
• Improved food supply
• Avoid competition with land-
• Degraded
based food production
• Low (Abolina and Luzadis, 2015; Baxter and
(through use of abandoned
Sustainable ecological Calvert, 2017; Calvert and Mabee, 2015;
lands)
Bioenergy & low-carbon value Campbell et al., 2008; Fargione et al.,
• Reduce emissions and
Renewables energy • Low potential 2008; Liu et al., 2017; Milbrandt et al.,
reliance on fossil fuels
production for natural 2014; Nijsen et al., 2012; Tilman et al.,
• Supplement domestic energy
regeneration/re 2006; Vuichard et al., 2009)
supplies
storation
• Rural economic opportunities

55
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

2.4.2 Soil carbon sequestration rates of proposals

Recognized for its role in climate change mitigation, soil carbon sequestration removes and
stores atmospheric carbon for longer periods of time compared to other terrestrial carbon pools
while providing valuable ecosystem co-benefits (Lal et al., 2015). Efforts to increase SOC
stocks of active agricultural lands, such as the international initiative “4 per 1000”, centre on
leveraging these features to offset anthropogenic greenhouse gas emissions (Minasny et al.,
2017). Abandoned agricultural lands therefore offer additional opportunities for large-scale
sequestration initiatives. Several global and biome-wide syntheses of LUC have shown that
biomass and SOC accumulation is most pronounced on lands previously cultivated that are
regenerating under passive or active management due to the tendency of these lands to be SOC
depleted (Deng et al., 2016; Don et al., 2011; Guo and Gifford, 2002; Kämpf et al., 2016;
Laganière et al., 2010; Li et al., 2018; Poeplau et al., 2011; Post and Kwon, 2000). In addition
to these studies, researchers have also reported soil carbon sequestration rates following
conventional agriculture conversion to low-intensity agriculture and bioenergy production.
This allows for a comparison of rates for these proposals synthesized over large geographic
scales (Figure 10).

Figure 10. Soil carbon sequestration rates of proposals implemented on abandoned or


converted agricultural lands synthesized at the country, climatic zone, and global scales.
Combined number of observations from studies of each proposal indicated in parentheses.

56
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

Numbers next to bars indicate studies in order from highest to lowest sequestration: 1Deng et
al., (2014), 11-30 years post-abandonment; 2Silver et al., (2000), 20 years post-abandonment;
3
Tilman et al., (2006), first 10 years, derived from net life cycle CO2 sequestration, includes
roots; 4Conant et al., (2001), cropland conversion to pasture; 5Feliciano et al., (2018),
cropland conversion to agrisilviculture; 6Li et al., (2012), mixed species plantations;
7
Kurganova et al., (2014), 20 years post-abandonment; 8Kämpf et al., (2016), various years
post-abandonment; 9Poeplau et al., (2011), 20 years post-abandonment; 10Wertebach et al.,
(2017), 20 years post-abandonment; 11Shi et al., (2013), 20 years post-abandonment;
12
McLauchlan et al., (2006), mixed species plantations; 13England et al., (2016), mixed species
plantations; 14Post and Kwon, (2000), various years post-abandonment; and 15Deng et al.,
(2016), cropland conversion to grassland.
All the proposals with reported sequestration rates had a positive impact on SOC over the first
two to three decades. The most effective approach is to combine passive rewilding and active
restoration techniques, enabling soil carbon sequestration rates as high as 1.30 Mg C ha−1 yr−1
across the tropical zone and in China (Deng et al., 2014; Silver et al., 2001). When synthesized
globally, however, active management through restoration alone has a higher maximum
reported sequestration potential compared to studies considering both active and passive
management techniques. By weighted average of all study observations, restoration at the
global scale sequestered 0.87 Mg C ha−1 yr−1 (Li et al., 2012; Shi et al., 2013) while restoration
combined with rewilding sequestered 0.34 Mg C ha−1 yr−1 (Deng et al., 2016; Post and Kwon,
2000).

Rewilding alone has higher reported rates than restoration at the country scale, although still
slightly lower than restoration at the global scale. The widespread ALA in the former Soviet
Union allowed for passive rewilding of 58 Mha in Russia and Kazakhstan (Kurganova et al.,
2015). The studies from this region included in Figure 10 estimate a weighted average
sequestration rate of 0.72 Mg C ha−1 yr−1 (Kurganova et al., 2014; Wertebach et al., 2017). This
exact rate has been observed across the entire temperate zone for grasslands regeneration post-
abandonment (Kämpf et al., 2016).

The high variability in rates reported between active, passive, and combined management
strategies is likely attributed to contrasting study parameters, including past (e.g., cropland vs.
pasture) and new (e.g., grassland vs. forest) land use, ecosystem characteristics, time since
ALA, geographic scale, sampling depth, and management approach and intensity. In Australia,
for example, tree plantings established on abandoned pastures in the northern wet tropics
resulted in no consistent soil carbon accumulation (Lewis et al., 2019), whereas southern and
eastern Australian plantations increased SOC stocks by 0.57 Mg C ha−1 yr−1 with more gains
on abandoned croplands than pastures (England et al., 2016).

57
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

Low-intensity agriculture results in substantial SOC accumulation, although with notably fewer
observations globally (Figure 10). The amount and quality of carbon input from aboveground
vegetation, both trees and crops, determines the rate and longevity of belowground
accumulation (Nair et al., 2009). For example, tree species diversity in agroforestry systems
implemented on previously cultivated lands and grasslands in India was positively linked to
increased sequestration, eventually producing near natural SOC levels (Hombegowda et al.,
2016).

Bioenergy production also features some of the highest rates of sequestration, in addition to
facilitating fossil fuel substitution. However, the carbon intensive production methods of many
sources of biofuel can limit their net positive impact. Only a few production methods, namely
grassland species grown specifically on abandoned lands, do not incur a carbon debt (Fargione
et al., 2008). When mixtures of grassland species are used, soil carbon sequestration can exceed
production emissions (Tilman et al., 2006) and outperform alternative management strategies
such as rewilding (Vuichard et al., 2009).

Considering the positive effect of combining active and passive management techniques
reported at some country and climatic zone scales, restoration initiatives may be most beneficial
for soil carbon sequestration when supplementing ongoing rewilding. Slow ecological
succession processes in degraded landscapes can be boosted from tree and shrub plantings (Rey
Benayas et al., 2008; Zethof et al., 2019). However, the wide range of sequestration rates
reported when combining restoration and rewilding around the world (0.30–1.30 Mg C ha−1
yr−1) illustrate the critical importance of site-specific factors in determining what approach or
combination of specific techniques to use. Even in cases where plantations clearly outperform
natural succession in terms of aboveground carbon sequestration, rewilding is still considered
an attractive option due to lower costs, higher plant biodiversity, and generally higher
belowground sequestration in the 0–30 cm depth (Tremblay and Ouimet, 2013).

These proposals have direct SOC implications. They need to be designed to promote their
above- and belowground accumulation potentials well beyond the first few decades (Vuichard
et al., 2009). Passive management and bioenergy production are also two of the most significant
climate change mitigation strategies to implement on active croplands directly (Albanito et al.,
2016). Policymakers providing mitigation incentives over large rural areas with or without
ALA should ensure landowners can reasonably consider one or more of these strategies based
on site-specific factors.

58
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

2.5 Conclusions
Establishing new land uses on abandoned agricultural lands is becoming increasingly attractive
as global demand for land, food, and energy intensifies. This presents notable opportunities for
carbon sequestration co-benefits. All six categories of proposals highlighted here have lasting
implications (directly or indirectly) for climate change mitigation efforts. From a soil carbon
sequestration perspective across large geographic scales, a combination of passive and active
management achieves the highest reported rates on abandoned lands.

Site-specific factors play a significant role in the efficacy of each proposal in each agricultural
region of the world, resulting in high variability among soil carbon sequestration rates. To
better quantify what past and present ALA implies for climate change mitigation, long-term
experiments featuring a variety of agricultural practices and crop types abandoned in different
bioclimates should be considered. This will require overcoming the challenge of gathering soil
data at the decadal scale and longer. Chronosequences of ALA and paired-plots substitutes
should be established extensively to supplement databases of observed sequestration rates. This
space-for-time substitution is a widely accepted method for determining the environmental
impacts of LUC over time (Walker et al., 2010).

The ecological and rural development implications of each management strategy and new land
use highlighted here informs policymakers tasked with planning the future of rural areas
experiencing ALA. Mitigating the detrimental impacts of climate change by returning carbon
to depleted soils will require exploring all available avenues, including leveraging abandoned
agricultural lands.

59
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

2.6 References
1. Abolina, E., Luzadis, V.A., 2015. Abandoned agricultural land and its potential for short rotation woody crops
in Latvia. Land use policy 49, 435–445. https://doi.org/10.1016/J.LANDUSEPOL.2015.08.022

2. Albanito, F., Beringer, T., Corstanje, R., Poulter, B., Stephenson, A., Zawadzka, J., Smith, P., 2016. Carbon
implications of converting cropland to bioenergy crops or forest for climate mitigation: A global assessment.
GCB Bioenergy 8, 81–95. https://doi.org/10.1111/gcbb.12242

3. Albert, Á.-J., Mudrák, O., Jongepierová, I., Fajmon, K., Frei, I., Ševčíková, M., Klimešová, J., Doležal, J.,
2019. Grassland restoration on ex-arable land by transfer of brush-harvested propagules and green hay. Agric.
Ecosyst. Environ. 272, 74–82. https://doi.org/10.1016/J.AGEE.2018.11.008

4. Arosa, M.L., Bastos, R., Cabral, J.A., Freitas, H., Costa, S.R., Santos, M., 2017. Long-term sustainability of
cork oak agro-forests in the Iberian Peninsula: A model-based approach aimed at supporting the best
management options for the montado conservation. Ecol. Modell. 343, 68–79.
https://doi.org/10.1016/j.ecolmodel.2016.10.008

5. Basualdo, M., Huykman, N., Volante, J.N., Paruelo, J.M., Piñeiro, G., 2019. Lost forever? Ecosystem
functional changes occurring after agricultural abandonment and forest recovery in the semiarid Chaco
forests. Sci. Total Environ. 650, 1537–1546. https://doi.org/10.1016/j.scitotenv.2018.09.001

6. Baumann, M., Kuemmerle, T., Elbakidze, M., Ozdogan, M., Radeloff, V.C., Keuler, N.S., Prishchepov, A.
V., Kruhlov, I., Hostert, P., 2011. Patterns and drivers of post-socialist farmland abandonment in Western
Ukraine. Land use policy 28, 552–562. https://doi.org/10.1016/j.landusepol.2010.11.003

7. Baxter, R.E., Calvert, K.E., 2017. Estimating Available Abandoned Cropland in the United States:
Possibilities for Energy Crop Production. Ann. Am. Assoc. Geogr. 107, 1162–1178.
https://doi.org/10.1080/24694452.2017.1298985

8. Benjamin, K., Bouchard, A., Domon, G., 2007. Abandoned farmlands as components of rural landscapes: An
analysis of perceptions and representations. Landsc. Urban Plan. 83, 228–244.
https://doi.org/10.1016/j.landurbplan.2007.04.009

9. Bryngelsson, D.K., Lindgren, K., 2013. Why large-scale bioenergy production on marginal land is unfeasible:
A conceptual partial equilibrium analysis. Energy Policy 55, 454–466.
https://doi.org/10.1016/J.ENPOL.2012.12.036

10. Calvert, K., Mabee, W., 2015. More solar farms or more bioenergy crops? Mapping and assessing potential
land-use conflicts among renewable energy technologies in eastern Ontario, Canada. Appl. Geogr. 56, 209–
221. https://doi.org/10.1016/J.APGEOG.2014.11.028

11. Campbell, J.E., Lobell, D.B., Genova, R.C., Field, C.B., 2008. The global potential of bioenergy on
abandoned agriculture lands. Environ. Sci. Technol. 42, 5791–5794. https://doi.org/10.1021/es800052w

12. Campbell, J.E., Lobell, D.B., Genova, R.C., Zumkehr, A., Gikonyo, C.B., 2013. Seasonal energy storage
using bioenergy production from abandoned croplands. Environ. Res. Lett. 8, 1–7.

13. Cerqueira, Y., Navarro, L.M., Maes, J., Marta-Pedroso, C., Honrado, J.P., Pereira, H.M., 2015. Ecosystem
services: The opportunities of rewilding in Europe, in: Rewilding European Landscapes. Springer
International Publishing, pp. 47–64. https://doi.org/10.1007/978-3-319-12039-3_3

14. Chapman, C.A., Chapman, L.J., 1999. Forest Restoration in Abandoned Agricultural Land: a Case Study
from East Africa, Conservation Biology.

15. Chen, B., Qui, Z., Nakamura, K., 2016. Tourist preferences for agricultural landscapes: a case study of
terraced paddy fields in Noto Peninsula, Japan. J. Mt. Sci. 13, 1880–1892. https://doi.org/10.1007/s11629-
015-3564-0

16. Conant, R.T., Paustian, K., Elliott, E.T., 2001. GRASSLAND MANAGEMENT AND CONVERSION INTO
GRASSLAND: EFFECTS ON SOIL CARBON, Ecological Applications.

60
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

17. Cramer, V.A., Hobbs, R.J., Standish, R.J., 2008. What’s new about old fields? Land abandonment and
ecosystem assembly. Trends Ecol. Evol. https://doi.org/10.1016/j.tree.2007.10.005

18. Crouzeilles, R., Ferreira, M.S., Chazdon, R.L., Lindenmayer, D.B., B Sansevero, J.B., Monteiro, L.,
Iribarrem, A., Latawiec, A.E., N Strassburg, B.B., 2017. Ecological restoration success is higher for natural
regeneration than for active restoration in tropical forests. Sci. Adv. 3.

19. De Beenhouwer, M., Aerts, R., Honnay, O., 2013. A global meta-analysis of the biodiversity and ecosystem
service benefits of coffee and cacao agroforestry. Agric. Ecosyst. Environ. 175, 1–7.
https://doi.org/10.1016/J.AGEE.2013.05.003

20. Deng, L., Liu, G. bin, Shangguan, Z. ping, 2014. Land-use conversion and changing soil carbon stocks in
China’s “Grain-for-Green” Program: A synthesis. Glob. Chang. Biol. https://doi.org/10.1111/gcb.12508

21. Deng, L., Zhu, G., Tang, Z., Shangguan, Z., 2016. Global patterns of the effects of land-use changes on soil
carbon stocks. Glob. Ecol. Conserv. 5, 127–138. https://doi.org/10.1016/J.GECCO.2015.12.004

22. Díaz, G.I., Nahuelhual, L., Echeverría, C., Marín, S., 2011. Drivers of land abandonment in Southern Chile
and implications for landscape planning. Landsc. Urban Plan. 99, 207–217.
https://doi.org/10.1016/j.landurbplan.2010.11.005

23. Don, A., Schumacher, J., Freibauer, A., 2011. Impact of tropical land-use change on soil organic carbon
stocks - a meta-analysis. Glob. Chang. Biol. https://doi.org/10.1111/j.1365-2486.2010.02336.x

24. Edrisi, S.A., Abhilash, P.C., 2016. Exploring marginal and degraded lands for biomass and bioenergy
production: An Indian scenario. Renew. Sustain. Energy Rev. 54, 1537–1551.
https://doi.org/10.1016/J.RSER.2015.10.050

25. England, J.R., Paul, K.I., Cunningham, S.C., Madhavan, D.B., Baker, T.G., Read, Z., Wilson, B.R.,
Cavagnaro, T.R., Lewis, T., Perring, M.P., Herrmann, T., Polglase, P.J., 2016. Previous land use and climate
influence differences in soil organic carbon following reforestation of agricultural land with mixed-species
plantings. Agric. Ecosyst. Environ. 227, 61–72. https://doi.org/10.1016/J.AGEE.2016.04.026

26. Estel, S., Kuemmerle, T., Alcántara, C., Levers, C., Prishchepov, A., Hostert, P., 2015. Mapping farmland
abandonment and recultivation across Europe using MODIS NDVI time series. Remote Sens. Environ. 163,
312–325. https://doi.org/10.1016/j.rse.2015.03.028

27. Fargione, J., Hill, J., Tilman, D., Polasky, S., Hawthorne, P., 2008. Land Clearing and the Biofuel Carbon
Debt. Science (80-. ). 319, 1235–1238. https://doi.org/10.1126/science.1153445

28. Feliciano, D., Ledo, A., Hillier, J., Nayak, D.R., 2018. Which agroforestry options give the greatest soil and
above ground carbon benefits in different world regions? Agric. Ecosyst. Environ. 254, 117–129.
https://doi.org/10.1016/J.AGEE.2017.11.032

29. García-Frapolli, E., Ayala-Orozco, B., Bonilla-Moheno, M., Espadas-Manrique, C., Ramos-Fernández, G.,
2007. Biodiversity conservation, traditional agriculture and ecotourism: Land cover/land use change
projections for a natural protected area in the northeastern Yucatan Peninsula, Mexico. Landsc. Urban Plan.
83, 137–153. https://doi.org/10.1016/J.LANDURBPLAN.2007.03.007

30. Garrido, P., Elbakidze, M., Angelstam, P., Plieninger, T., Pulido, F., Moreno, G., 2017. Stakeholder
perspectives of wood-pasture ecosystem services: A case study from Iberian dehesas. Land use policy 60,
324–333. https://doi.org/10.1016/j.landusepol.2016.10.022

31. Griffiths, P., Müller, D., Kuemmerle, T., Hostert, P., 2013. Agricultural land change in the Carpathian
ecoregion after the breakdown of socialism and expansion of the European Union. Environ. Res. Lett. 8.
https://doi.org/10.1088/1748-9326/8/4/045024

32. Guo, L.B., Gifford, R.M., 2002. Soil carbon stocks and land use change: A meta analysis. Glob. Chang. Biol.
https://doi.org/10.1046/j.1354-1013.2002.00486.x

61
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

33. Hall, S.J.G., 2018. A novel agroecosystem: Beef production in abandoned farmland as a multifunctional
alternative to rewilding. Agric. Syst. 167, 10–16. https://doi.org/10.1016/J.AGSY.2018.08.009

34. Hanaček, K., Rodríguez-Labajos, B., 2018. Impacts of land-use and management changes on cultural
agroecosystem services and environmental conflicts—A global review. Glob. Environ. Chang. 50, 41–59.
https://doi.org/10.1016/J.GLOENVCHA.2018.02.016

35. Heinimann, A., Mertz, O., Frolking, S., Christensen, A.E., Hurni, K., Sedano, F., Chini, L.P., Sahajpal, R.,
Hansen, M., Hurtt, G., 2017. A global view of shifting cultivation: Recent, current, and future extent. PLoS
One 12. https://doi.org/10.1371/journal.pone.0184479

36. Hombegowda, H.C., Van Straaten, O., Köhler, M., Hölscher, D., 2016. On the rebound: Soil organic carbon
stocks can bounce back to near forest levels when agroforests replace agriculture in southern India. SOIL.
https://doi.org/10.5194/soil-2-13-2016

37. IPCC, 2019. IPCC Special Report on Climate Change, Desertification, Land Degradation, Sustainable Land
Management, Food Security, and Greenhouse gas fluxes in Terrestrial Ecosystems. Summary for
Policymakers. Geneva, Switzerland.

38. Jezeer, R.E., Verweij, P.A., Santos, M.J., Boot, R.G.A., 2017. Shaded Coffee and Cocoa – Double Dividend
for Biodiversity and Small-scale Farmers. Ecol. Econ. 140, 136–145.
https://doi.org/10.1016/J.ECOLECON.2017.04.019

39. Jiao, F., Wen, Z.-M., An, S.-S., 2011. Changes in soil properties across a chronosequence of vegetation
restoration on the Loess Plateau of China. CATENA 86, 110–116.
https://doi.org/10.1016/J.CATENA.2011.03.001

40. Kämpf, I., Hölzel, N., Störrle, M., Broll, G., Kiehl, K., 2016. Potential of temperate agricultural soils for
carbon sequestration: A meta-analysis of land-use effects. Sci. Total Environ. 566–567, 428–435.
https://doi.org/10.1016/J.SCITOTENV.2016.05.067

41. Knapik, W., 2018. The innovative model of Community-based Social Farming (CSF). J. Rural Stud. 60, 93–
104. https://doi.org/10.1016/J.JRURSTUD.2018.03.008

42. Knoke, T., Bendix, J., Pohle, P., Hamer, U., Hildebrandt, P., Roos, K., Gerique, A., Sandoval, M.L., Breuer,
L., Tischer, A., Silva, B., Calvas, B., Aguirre, N., Castro, L.M., Windhorst, D., Weber, M., Stimm, B., Günter,
S., Palomeque, X., Mora, J., Mosandl, R., Beck, E., 2014. Afforestation or intense pasturing improve the
ecological and economic value of abandoned tropical farmlands. Nat. Commun. 5.
https://doi.org/10.1038/ncomms6612

43. Knoke, T., Calvas, B., Moreno, S.O., Onyekwelu, J.C., Griess, V.C., 2013. Food production and climate
protection—What abandoned lands can do to preserve natural forests. Glob. Environ. Chang. 23, 1064–1072.
https://doi.org/10.1016/J.GLOENVCHA.2013.07.004

44. Kurganova, I., Lopes de Gerenyu, V., Kuzyakov, Y., 2015. Large-scale carbon sequestration in post-
agrogenic ecosystems in Russia and Kazakhstan. CATENA 133, 461–466.
https://doi.org/10.1016/J.CATENA.2015.06.002

45. Kurganova, I., Lopes de Gerenyu, V., Six, J., Kuzyakov, Y., 2014. Carbon cost of collective farming collapse
in Russia. Glob. Chang. Biol. https://doi.org/10.1111/gcb.12379

46. Laganière, J., Angers, D.A., Paré, D., 2010. Carbon accumulation in agricultural soils after afforestation: A
meta-analysis. Glob. Chang. Biol. 16, 439–453. https://doi.org/10.1111/j.1365-2486.2009.01930.x

47. Lal, R., Negassa, W., Lorenz, K., 2015. Carbon sequestration in soil. Curr. Opin. Environ. Sustain.
https://doi.org/10.1016/j.cosust.2015.09.002

48. Lasanta, T., Nadal-Romero, E., Arnáez, J., 2015. Managing abandoned farmland to control the impact of re-
vegetation on the environment. The state of the art in Europe. Environ. Sci. Policy 52, 99–109.
https://doi.org/10.1016/J.ENVSCI.2015.05.012

62
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

49. Lewis, T., Verstraten, L., Hogg, B., Wehr, B.J., Swift, S., Tindale, N., Menzies, N.W., Dalal, R.C., Bryant,
P., Francis, B., Smith, T.E., 2019. Reforestation of agricultural land in the tropics: The relative contribution
of soil, living biomass and debris pools to carbon sequestration. Sci. Total Environ. 649, 1502–1513.
https://doi.org/10.1016/j.scitotenv.2018.08.351

50. Li, D., Niu, S., Luo, Y., 2012. Global patterns of the dynamics of soil carbon and nitrogen stocks following
afforestation: A meta-analysis. New Phytol. 195, 172–181. https://doi.org/10.1111/j.1469-
8137.2012.04150.x

51. Li, W., Ciais, P., Guenet, B., Peng, S., Chang, J., Chaplot, V., Khudyaev, S., Peregon, A., Piao, S., Wang, Y.,
Yue, C., 2018. Temporal response of soil organic carbon after grassland‐related land‐use change. Glob.
Chang. Biol. https://doi.org/10.1111/gcb.14328

52. Liu, T., Huffman, T., Kulshreshtha, S., McConkey, B., Du, Y., Green, M., Liu, J., Shang, J., Geng, X., 2017.
Bioenergy production on marginal land in Canada: Potential, economic feasibility, and greenhouse gas
emissions impacts. Appl. Energy 205, 477–485. https://doi.org/10.1016/J.APENERGY.2017.07.126

53. Londoño, A.C., Williams, P.R., Hart, M.L., 2017. A change in landscape: Lessons learned from abandonment
of ancient Wari agricultural terraces in Southern Peru. J. Environ. Manage. 202, 532–542.
https://doi.org/10.1016/j.jenvman.2017.01.012

54. Mander, Ü., Mikk, M., Külvik, M., 1999. Ecological and low intensity agriculture as contributors to landscape
and biological diversity. Landsc. Urban Plan. 46, 169–177. https://doi.org/10.1016/S0169-2046(99)00042-0

55. McLauchlan, K.K., Hobbie, S.E., Post, W.M., 2006. Conversion from agriculture to grassland builds soil
organic matter on decadal timescales. Ecol. Appl. 16, 143–153. https://doi.org/10.1890/04-1650

56. Meli, P., Holl, K.D., Benayas, J.M.R., Jones, H.P., Jones, P.C., Montoya, D., Mateos, D.M., 2017. A global
review of past land use, climate, and active vs. passive restoration effects on forest recovery. PLoS One.
https://doi.org/10.1371/journal.pone.0171368

57. Merckx, T., Pereira, H.M., 2015. Reshaping agri-environmental subsidies: From marginal farming to large-
scale rewilding. Basic Appl. Ecol. 16, 95–103. https://doi.org/10.1016/J.BAAE.2014.12.003

58. Milbrandt, A.R., Heimiller, D.M., Perry, A.D., Field, C.B., 2014. Renewable energy potential on marginal
lands in the United States. Renew. Sustain. Energy Rev. 29, 473–481.
https://doi.org/10.1016/J.RSER.2013.08.079

59. Minasny, B., Malone, B.P., McBratney, A.B., Angers, D.A., Arrouays, D., Chambers, A., Chaplot, V., Chen,
Z.S., Cheng, K., Das, B.S., Field, D.J., Gimona, A., Hedley, C.B., Hong, S.Y., Mandal, B., Marchant, B.P.,
Martin, M., McConkey, B.G., Mulder, V.L., O’Rourke, S., Richer-de-Forges, A.C., Odeh, I., Padarian, J.,
Paustian, K., Pan, G., Poggio, L., Savin, I., Stolbovoy, V., Stockmann, U., Sulaeman, Y., Tsui, C.C.,
V�gen, T.G., van Wesemael, B., Winowiecki, L., 2017. Soil carbon 4 per mille. Geoderma.
https://doi.org/10.1016/j.geoderma.2017.01.002

60. Morimoto, Y., 2011. What is Satoyama? Points for discussion on its future direction. Landsc. Ecol. Eng. 7,
163–171. https://doi.org/10.1007/s11355-010-0120-5

61. Nair, P.K.R., Kumar, B.M., Nair, V.D., 2009. Agroforestry as a strategy for carbon sequestration. J. Plant
Nutr. Soil Sci. https://doi.org/10.1002/jpln.200800030

62. Navarro, L.M., Pereira, H.M., 2012. Rewilding abandoned landscapes in Europe, in: Rewilding European
Landscapes. Springer International Publishing, pp. 3–23. https://doi.org/10.1007/978-3-319-12039-3_1

63. Nijsen, M., Smeets, E., Stehfest, E., van Vuuren, D.P., 2012. An evaluation of the global potential of
bioenergy production on degraded lands. GCB Bioenergy 4, 130–147. https://doi.org/10.1111/j.1757-
1707.2011.01121.x

64. Osawa, T., Kohyama, K., Mitsuhashi, H., 2016. Multiple factors drive regional agricultural abandonment.
Sci. Total Environ. 542, 478–483. https://doi.org/10.1016/J.SCITOTENV.2015.10.067

63
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

65. Pace Ricci, J.M., Conrad, E., 2018. Exploring the feasibility of setting up community allotments on
abandoned agricultural land: A place, people, policy approach. Land use policy 79, 102–115.
https://doi.org/10.1016/J.LANDUSEPOL.2018.08.009

66. Pelorosso, R., Della Chiesa, S., Tappeiner, U., Leone, A., Rocchini, D., 2011. Stability analysis for defining
management strategies in abandoned mountain landscapes of the Mediterranean basin. Landsc. Urban Plan.
103, 335–346. https://doi.org/10.1016/J.LANDURBPLAN.2011.08.007

67. Perino, A., Pereira, H.M., Navarro, L.M., Fernández, N., Bullock, J.M., Ceauşu, S., Cortés-Avizanda, A.,
Van Klink, R., Kuemmerle, T., Lomba, A., Pe’er, G., Plieninger, T., Benayas, J.M.R., Sandom, C.J.,
Svenning, J.C., Wheeler, H.C., 2019. Rewilding complex ecosystems. Science (80-. ).
https://doi.org/10.1126/science.aav5570

68. Plieninger, T., Hui, C., Gaertner, M., Huntsinger, L., 2014. The impact of land abandonment on species
richness and abundance in the Mediterranean Basin: A meta-analysis. PLoS One.
https://doi.org/10.1371/journal.pone.0098355

69. Poeplau, C., Don, A., Vesterdal, L., Leifeld, J., Van Wesemael, B., Schumacher, J., Gensior, A., 2011.
Temporal dynamics of soil organic carbon after land-use change in the temperate zone - carbon response
functions as a model approach. Glob. Chang. Biol. https://doi.org/10.1111/j.1365-2486.2011.02408.x

70. Post, W.M., Kwon, K.C., 2000. Soil carbon sequestration and land‐use change: processes and potential. Glob.
Chang. Biol. 6. https://doi.org/10.1046/j.1365-2486.2000.00308.x

71. Prach, K., Jongepierová, I., Řehounková, K., Fajmon, K., 2014. Restoration of grasslands on ex-arable land
using regional and commercial seed mixtures and spontaneous succession: Successional trajectories and
changes in species richness. Agric. Ecosyst. Environ. 182, 131–136.
https://doi.org/10.1016/J.AGEE.2013.06.003

72. Queiroz, C., Beilin, R., Folke, C., Lindborg, R., 2014. Farmland abandonment: Threat or opportunity for
biodiversity conservation? A global review. Front. Ecol. Environ. https://doi.org/10.1890/120348

73. Regos, A., Domínguez, J., Gil-Tena, A., Brotons, L., Ninyerola, M., Pons, X., 2016. Rural abandoned
landscapes and bird assemblages: winners and losers in the rewilding of a marginal mountain area (NW
Spain). Reg. Environ. Chang. 16, 199–211. https://doi.org/10.1007/s10113-014-0740-7

74. Rey Benayas, J.M., Bullock, J.M., Newton, A.C., 2008. Creating woodland islets to reconcile ecological
restoration, conservation, and agricultural land use. Front. Ecol. Environ. https://doi.org/10.1890/070057

75. Rey Benayas, J.M., Martins, A., Nicolau, J.M., Schulz, J.J., 2007. Abandonment of agricultural land: an
overview of drivers and consequences. CAB Rev. Perspect. Agric. Vet. Sci. Nutr. Nat. Resour.
https://doi.org/10.1079/PAVSNNR20072057

76. Robledano-Aymerich, F., Romero-Díaz, A., Belmonte-Serrato, F., Zapata-Pérez, V.M., Martínez-Hernández,
C., Martínez-López, V., 2014. Ecogeomorphological consequences of land abandonment in semiarid
Mediterranean areas: Integrated assessment of physical evolution and biodiversity. Agric. Ecosyst. Environ.
197, 222–242. https://doi.org/10.1016/J.AGEE.2014.08.006

77. Rocha, G.P.E., Vieira, D.L.M., Simon, M.F., 2016. Fast natural regeneration in abandoned pastures in
southern Amazonia. For. Ecol. Manage. 370, 93–101. https://doi.org/10.1016/J.FORECO.2016.03.057

78. Ruskule, A., Nikodemus, O., Kasparinskis, R., Bell, S., Urtane, I., 2013. The perception of abandoned
farmland by local people and experts: Landscape value and perspectives on future land use. Landsc. Urban
Plan. 115, 49–61. https://doi.org/10.1016/j.landurbplan.2013.03.012

79. Sanagustín-Fons, M.V., Fierro, J.A.M., Patiño, M.G. y, 2011. Rural tourism: A sustainable alternative. Appl.
Energy 88, 551–557. https://doi.org/10.1016/J.APENERGY.2010.08.031

80. Sansevero, J.B.B., Prieto, P.V., Sánchez-Tapia, A., Braga, J.M.A., Rodrigues, P.J.F.P., 2017. Past land-use
and ecological resilience in a lowland Brazilian Atlantic Forest: implications for passive restoration. New
For. 48, 573–586. https://doi.org/10.1007/s11056-017-9586-4

64
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

81. Santos, P.Z.F., Crouzeilles, R., Sansevero, J.B.B., 2019. Can agroforestry systems enhance biodiversity and
ecosystem service provision in agricultural landscapes? A meta-analysis for the Brazilian Atlantic Forest.
For. Ecol. Manage. 433, 140–145. https://doi.org/10.1016/J.FORECO.2018.10.064

82. Sayadi, S., González-Roa, M.C., Calatrava-Requena, J., 2009. Public preferences for landscape features: The
case of agricultural landscape in mountainous Mediterranean areas. Land use policy 26, 334–344.
https://doi.org/10.1016/j.landusepol.2008.04.003

83. Schierhorn, F., Müller, D., Beringer, T., Prishchepov, A. V., Kuemmerle, T., Balmann, A., 2013. Post-Soviet
cropland abandonment and carbon sequestration in European Russia, Ukraine, and Belarus. Global
Biogeochem. Cycles. https://doi.org/10.1002/2013GB004654

84. Schröder, P., Beckers, B., Daniels, S., Gnädinger, F., Maestri, E., Marmiroli, N., Mench, M., Millan, R.,
Obermeier, M.M., Oustriere, N., Persson, T., Poschenrieder, C., Rineau, F., Rutkowska, B., Schmid, T.,
Szulc, W., Witters, N., Sæbø, A., 2018. Intensify production, transform biomass to energy and novel goods
and protect soils in Europe—A vision how to mobilize marginal lands. Sci. Total Environ. 616–617, 1101–
1123. https://doi.org/10.1016/J.SCITOTENV.2017.10.209

85. Schulp, C.J.E., Levers, C., Kuemmerle, T., Tieskens, K.F., Verburg, P.H., 2018. Mapping and modelling past
and future land use change in Europe’s cultural landscapes. Land use policy.
https://doi.org/10.1016/j.landusepol.2018.04.030

86. Scott, A.J., Morgan, J.W., 2012. Recovery of soil and vegetation in semi-arid Australian old fields. J. Arid
Environ. 76, 61–71. https://doi.org/10.1016/J.JARIDENV.2011.08.014

87. Shahid, S.A., Al-Shankiti, A., 2013. Sustainable food production in marginal lands—Case of GDLA member
countries. Int. Soil Water Conserv. Res. 1, 24–38. https://doi.org/10.1016/S2095-6339(15)30047-2

88. Shi, S., Zhang, W., Zhang, P., Yu, Y., Ding, F., 2013. A synthesis of change in deep soil organic carbon stores
with afforestation of agricultural soils. For. Ecol. Manage. 296, 53–63.
https://doi.org/10.1016/J.FORECO.2013.01.026

89. Shortall, O.K., 2013. “Marginal land” for energy crops: Exploring definitions and embedded assumptions.
Energy Policy 62, 19–27. https://doi.org/10.1016/J.ENPOL.2013.07.048

90. Silver, W.L., Ostertag, R., Lugo, A.E., 2000. The Potential for Carbon Sequestration Through Reforestation
of Abandoned Tropical Agricultural and Pasture Lands.

91. Smaliychuk, A., Müller, D., Prishchepov, A. V., Levers, C., Kruhlov, I., Kuemmerle, T., 2016. Recultivation
of abandoned agricultural lands in Ukraine: Patterns and drivers. Glob. Environ. Chang. 38, 70–81.
https://doi.org/10.1016/J.GLOENVCHA.2016.02.009

92. Smith, J., Pearce, B.D., Wolfe, M.S., 2013. Reconciling productivity with protection of the environment: Is
temperate agroforestry the answer? Renew. Agric. Food Syst. https://doi.org/10.1017/S1742170511000585

93. Sojneková, M., Chytrý, M., 2015. From arable land to species-rich semi-natural grasslands: Succession in
abandoned fields in a dry region of central Europe. Ecol. Eng. 77, 373–381.
https://doi.org/10.1016/J.ECOLENG.2015.01.042

94. Swinton, S.M., Babcock, B.A., James, L.K., Bandaru, V., 2011. Higher US crop prices trigger little area
expansion so marginal land for biofuel crops is limited. Energy Policy 39, 5254–5258.
https://doi.org/10.1016/J.ENPOL.2011.05.039

95. Tilman, D., Hill, J., Lehman, C., 2006. Carbon-Negative Biofuels from Low-Input High-Diversity Grassland
Biomass. Science (80-. ). 314, 1598–1600. https://doi.org/10.1126/science.1133306

96. Tomaz, C., Alegria, C., Monteiro, J.M., Teixeira, M.C., 2013. Land cover change and afforestation of
marginal and abandoned agricultural land: A 10year analysis in a Mediterranean region. For. Ecol. Manage.
308, 40–49. https://doi.org/10.1016/j.foreco.2013.07.044

65
CHAPTER II: Management opportunities for soil carbon sequestration following agricultural land abandonment

97. Torralba, M., Fagerholm, N., Burgess, P.J., Moreno, G., Plieninger, T., 2016. Do European agroforestry
systems enhance biodiversity and ecosystem services? A meta-analysis. Agric. Ecosyst. Environ. 230, 150–
161. https://doi.org/10.1016/J.AGEE.2016.06.002

98. Tremblay, S., Ouimet, R., 2013. White Spruce Plantations on Abandoned Agricultural Land: Are They More
Effective as C Sinks than Natural Succession? Forests. https://doi.org/10.3390/f4041141

99. Tscharntke, T., Clough, Y., Bhagwat, S.A., Buchori, D., Faust, H., Hertel, D., Hölscher, D., Juhrbandt, J.,
Kessler, M., Perfecto, I., Scherber, C., Schroth, G., Veldkamp, E., Wanger, T.C., 2011. Multifunctional shade-
tree management in tropical agroforestry landscapes - A review. J. Appl. Ecol. https://doi.org/10.1111/j.1365-
2664.2010.01939.x

100. Valdivia, C., Barbieri, C., 2014. Agritourism as a sustainable adaptation strategy to climate change in the
Andean Altiplano. Tour. Manag. Perspect. 11, 18–25. https://doi.org/10.1016/j.tmp.2014.02.004

101. Valentine, J., Clifton-Brown, J., Hastings, A., Robson, P., Allison, G., Smith, P., 2012. Food vs. fuel: The
use of land for lignocellulosic “next generation” energy crops that minimize competition with primary food
production. GCB Bioenergy 4, 1–19. https://doi.org/10.1111/j.1757-1707.2011.01111.x

102. van der Zanden, E.H., Carvalho-Ribeiro, S.M., Verburg, P.H., 2018. Abandonment landscapes: user attitudes,
alternative futures and land management in Castro Laboreiro, Portugal. Reg. Environ. Chang.
https://doi.org/10.1007/s10113-018-1294-x

103. Varotto, M., Lodatti, L., 2014. New Family Farmers for Abandoned Lands. Mt. Res. Dev. 34, 315–325.
https://doi.org/10.1659/mrd-journal-d-14-00012.1

104. Verburg, P.H., Overmars, K.P., 2009. Combining top-down and bottom-up dynamics in land use modeling:
Exploring the future of abandoned farmlands in Europe with the Dyna-CLUE model. Landsc. Ecol. 24, 1167–
1181. https://doi.org/10.1007/s10980-009-9355-7

105. Vuichard, N., Ciais, P., Wolf, A., 2009. Soil carbon sequestration or biofuel production: New land-use
opportunities for mitigating climate over abandoned soviet farmlands. Environ. Sci. Technol. 43, 8678–8683.
https://doi.org/10.1021/es901652t

106. Walker, L.R., Wardle, D.A., Bardgett, R.D., Clarkson, B.D., 2010. The use of chronosequences in studies of
ecological succession and soil development. J. Ecol. https://doi.org/10.1111/j.1365-2745.2010.01664.x

107. Wang, L., Pijanowski, B., Yang, W., Zhai, R., Omrani, H., Li, K., 2018. Predicting multiple land use
transitions under rapid urbanization and implications for land management and urban planning: The case of
Zhanggong District in central China. Habitat Int. 82, 48–61.
https://doi.org/10.1016/J.HABITATINT.2018.08.007

108. Wertebach, T.M., Hölzel, N., Kämpf, I., Yurtaev, A., Tupitsin, S., Kiehl, K., Kamp, J., Kleinebecker, T.,
2017. Soil carbon sequestration due to post-Soviet cropland abandonment: estimates from a large-scale soil
organic carbon field inventory. Glob. Chang. Biol. 23, 3729–3741. https://doi.org/10.1111/gcb.13650

109. Yin, H., Prishchepov, A. V., Kuemmerle, T., Bleyhl, B., Buchner, J., Radeloff, V.C., 2018. Mapping
agricultural land abandonment from spatial and temporal segmentation of Landsat time series. Remote Sens.
Environ. 210, 12–24. https://doi.org/10.1016/j.rse.2018.02.050

110. Zagaria, C., Schulp, C.J.E., Kizos, T., Verburg, P.H., 2018. Perspectives of farmers and tourists on
agricultural abandonment in east Lesvos, Greece. Reg. Environ. Chang. 1–13.
https://doi.org/10.1007/s10113-017-1276-4

111. Zethof, J.H.T., Cammeraat, E.L.H., Nadal-Romero, E., 2019. The enhancing effect of afforestation over
secondary succession on soil quality under semiarid climate conditions. Sci. Total Environ. 652, 1090–1101.
https://doi.org/10.1016/J.SCITOTENV.2018.10.235

112. Zhang, Y., Li, X., Song, W., 2014. Determinants of cropland abandonment at the parcel, household and
village levels in mountain areas of China: A multi-level analysis. Land use policy 41, 186–192.
https://doi.org/10.1016/J.LANDUSEPOL.2014.05.011

66
CHAPTER III: Soil organic carbon accumulation
rates on Mediterranean abandoned agricultural
lands
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

3.1 Overview
Secondary succession on abandoned agricultural lands can produce climate change mitigation
co-benefits, such as soil carbon sequestration. However, the accumulation of soil organic
carbon (SOC) in Mediterranean regions has been difficult to predict and is subject to multiple
environmental and land management factors. Gains, losses, and no significant changes have all
been reported. Here, I compiled chronosequence data (n = 113) from published studies and new
field sites to assess the response of SOC to agricultural land abandonment in peninsular Spain.
I found an overall SOC concentration accumulation rate of +2.3% yr–1 post-abandonment. SOC
dynamics are highly variable and context-dependent. Minimal change occurs on abandoned
cereal croplands compared to abandoned woody croplands (+4% yr–1). Accumulation is most
prevalent within a Goldilocks climatic window of ~13–17 ° C and ~450–900 mm precipitation,
promoting > 100% gains after three decades. Our secondary forest field sites accrued 40.8 Mg
C ha–1 (+172%) following abandonment and displayed greater SOC and N depth heterogeneity
than natural forests demonstrating the long-lasting impact of agriculture. Although changes in
regional climate and crop types abandoned will impact future carbon sequestration,
abandonment remains a low-cost, long-term natural climate solution best incorporated in
tandem with other multipurpose sustainable land management strategies.

68
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

3.2 Introduction
Agricultural land abandonment (ALA) has been a part of Mediterranean landscape dynamics
for millennia (Butzer, 2005). Increases in key drivers such as agricultural intensification, rural
depopulation, and soil degradation have fuelled surges of widespread abandonment over the
course of the last century (Lasanta et al., 2017; Rey Benayas et al., 2007; Ustaoglu and Collier,
2018). While ALA is often associated with negative economic and sociocultural impacts for
landowners, decision-makers and rural tourists alike (Benjamin et al., 2007; Faccioni et al.,
2019; Ruskule et al., 2013; van der Zanden et al., 2018), the potential for Mediterranean ex-
arable lands to act as valuable carbon sinks though above- and belowground carbon
accumulation during ecological succession has become a research and policy focus (Chiti et
al., 2018; Novara et al., 2017; Pellis et al., 2019; Vilà-Cabrera et al., 2017).

Spain is one of the most vulnerable countries to ALA; its agricultural land area has contracted
by a fifth over the last six decades (FAO, 2019). Compared to the European Union average of
three percent, five percent of Spanish agricultural lands are projected to be abandoned by 2030
with over five million hectares (ha) at severe risk of abandonment (Castillo et al., 2020).
Mountainous agricultural landscapes in Spain have commonly experienced over 50%
abandonment during the 20th century (Lasanta et al., 2017), with some catchments reaching
near total (> 95%) reductions (Arnaez et al., 2011; Cohen et al., 2011). Considering such
extensive historical and projected coverage, understanding the ecosystem implications of ALA
is crucial for supporting rural development planning and improving national carbon inventories
(Padilla et al., 2010; Pausas and Millán, 2019).

Conventional agricultural practices are known to both deplete soil organic carbon (SOC) stocks
over time and emit substantial quantities of greenhouse gases (Carlson et al., 2017; Lal, 2013).
In contrast, the spontaneous recovery of SOC following the cessation of cultivation has been
identified as a potential avenue for climate change mitigation through soil carbon sequestration
(SCS) (Kämpf et al., 2016; Kurganova et al., 2014; Laganière et al., 2010; Wertebach et al.,
2017). While the SCS potential of most agricultural soils transitioning to grasslands,
shrublands, and forests has been established (Deng et al., 2014a; Guo and Gifford, 2002; Post
and Kwon, 2000), its generality as a universal and guaranteed response to abandonment
worldwide is unclear (Breuer et al., 2006; Hoogmoed et al., 2012). In semi-arid landscapes in
particular, intensive agriculture can often degrade soils that then require much longer periods
of natural or assisted restoration to overcome stalled vegetation recovery (Bonet, 2004; Garcia-
Franco et al., 2014; Robledano-Aymerich et al., 2014).

69
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

Several authors have observed slow or negligible SOC accumulation post-abandonment across
Spain: in the southeast under semi-arid conditions (Lesschen et al., 2008; Ruiz-Navarro et al.,
2009; Segura et al., 2020; Zethof et al., 2019); in the central restrictive gypsum steppes
(Eugenio et al., 2012; Martinez-Duro et al., 2010); and in the northeastern humid Pyrenees
(Nadal-Romero et al., 2016; Navas et al., 2012). However, there have also been reports of
original SOC concentrations increasing substantially on abandoned orchards and vineyards in
both northeastern (Dunjó et al., 2003; Emran et al., 2012; Pardini et al., 2003) and southeastern
Spain (De Baets et al., 2012; Rodrigo-Comino et al., 2018; Romero-Díaz et al., 2017), with
similar results observed in northern (Chiti et al., 2018) and southern Italy (Badalamenti et al.,
2019; Novara et al., 2013). This dichotomy in SOC response during Mediterranean post-
agricultural secondary succession presents challenges for decision-makers involved in the
management of regions undergoing ALA. It also demonstrates the need for increased research
efforts to clarify the key factors driving positive or negative responses.

Here, I compiled chronosequence data from published and new field sites to assess the response
of SOC to ALA in peninsular Spain. After first reviewing published study sites, I sampled three
new chronosequence locations in an underrepresented region of northeastern Spain to
investigate changes in SOC and soil N stocks at different stages of secondary succession and
soil depths. I then compared these field results with published data to determine if ALA has
generally led to increases or decreases in SOC concentrations across a range of Mediterranean
environments. In doing so, I explored the drivers modulating the rate of SOC gain or loss over
time. I hypothesize increasing SOC concentrations during secondary succession on average.
This paper aims to shed new light on the debate surrounding the long-term SCS value of
Mediterranean ALA.

3.3 Methods
3.3.1 Field sites
Three new chronosequences featuring six field sites of abandoned olive groves and vineyards
and six control sites were selected in Barcelona Province of Catalonia, Spain (NE of the Iberian
Peninsula), hereafter referred to as Font-rubí, Torrelavit, and Cànoves (Table 4). In this region,
just over six percent of all agricultural land is currently classified as abandoned, with ALA and
reforestation among the four most important land pressures (Baśnou et al., 2013; Funes et al.,
2019). All the field sites fall within a transitional zone between Csa (Temperate, dry and hot
summer) and Cfb (Temperate, no dry season, hot summer) Köppen–Geiger climates (Beck et
al., 2018), and are located within the Catalan Coastal Depression between the Mediterranean

70
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

Sea and the Catalan Pre-Coastal Range. The sites share similar underlying lithologies of
sediments and sedimentary rocks originating in the Neogene Period. Font-rubí and Torrelavit
are both located within the Penedès wine producing region while Cànoves is located in the
Vallès Oriental region just south of the Montseny Massif, a UNESCO Biosphere reserve. Their
soils are characterized by high calcium carbonate contents but support a diverse assemblage of
herb and shrub vegetation. Tree species dominating the forested sites include Pinus halepensis,
Quercus ilex, and Quercus cerrioides, with an occasional presence of Pinus pinea, Arbutus
unedo and Rhamnus alaternus. Occasional legacy growth of Olea europaea (European olive)
and Vitis vinifera (common grapevine) is also present (Figure 11.c). Management of olive
groves and vineyards in this region involves periodic ploughing, branch pruning and removal,
and moderate manure, herbicide, and mineral fertilizer application (García et al., 2007).

Table 4. Geographical characteristics and soil classification of the chronosequence field sites.

Soil
Soil Moisture
MAP MAT e.a.s.l. Soil Type Temperature
Chronosequence Location (USDA Soil
(mm) (° C) (m) (WRB, 2014) (USDA Soil
Taxonomy)
Taxonomy)

Calcaric
41°25'35.5"N, 420– Leptosols and
Font-rubí 650 15.0 Thermic Xeric
1°36'33.5"E 500 Petric
Calcisols

Calcaric
41°26'07.5"N, 250– Leptosols and
Torrelavit 600 15.5 Thermic Xeric
1°43'49.4"E 320 Petric
Calcisols

Calcaric
41°40'55.4"N, 300– Cambisols
Cànoves 696 15.1 Thermic Ustic
2°18'52.3"E 370 and Haplic
Calcisols

MAP: mean annual precipitation; MAT: mean annual temperature; e.a.s.l.: elevation above sea level.

71
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

3.3.1.1 Site selection and sampling


Near each location in Table 4, one natural control (late-stage forest), one pre-abandonment
control presently under cultivation (cropland), and two secondary forests previously cultivated
(early- and mid-stage forests) were selected to create three chronosequences of ALA (Figure
11). Our chronosequences therefore do not include stalled vegetation recovery, nor grassland
or shrubland landcovers. Historical orthophotography databases (SignA, Spanish National
Geographic Institute) and Landsat imagery (SatCat, Catalonia Satellite Image Server, CREAF-
UAB) were compared at different time intervals to determine the approximate date of
abandonment and if the subsequent secondary forest development was uninterrupted until
present (Breuer et al., 2006; Gabarrón-Galeote et al., 2015b; García et al., 2007; Lesschen et
al., 2008). Late-stage forests were classified based on having continuous forest coverage since
at least 1956 to represent soils that were presumably never cultivated. Further age verification
was conducted by consulting landowners of selected or adjoining properties and through
ground truthing during field visits. In each chronosequence, at least two of the four fields were
directly adjoining and all fields were within a 1.5 km radius to ensure similar environmental
conditions. Sampling was undertaken in spring 2019. As all plots were < 5 ha, three sampling
sites were chosen 50 m apart from each other based on similar aspect and slope in each
successional stage of each chronosequence (Stolbovoy et al., 2007). To reduce any affect of
field boundaries, a distance of approximately 50 m was also maintained between sampling sites
and the borders of other land covers. The sampling sites featured generally flat, shallow soils
approximately 30–50 cm deep. At each sampling site, the litter layer was removed and a trench
was dug in 10 cm increments to a depth of 30 cm. This depth was chosen according to IPCC
guidelines for the determination of SOC inventories following land use change (IPCC, 2006),
however ALA can also have notable impacts deeper in the soil profile especially after several
decades (Beniston et al., 2014). At each increment, 500 g of soil was sampled resulting in a
total of 108 samples. Undisturbed samples were also collected at each sampling site using a
100 cm3 steel ring for bulk density determination.

72
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

Figure 11. Orthophoto comparison of the Font-rubí chronosequence between 2019 (upper left)
and 1956 (lower left): (a) active cropland; (b) early-stage forest; (c) mid-stage forest; and (d)
late-stage forest. Ground evidence of past agricultural land use includes remnants of stone
terrace walls and abandoned dry stone huts (b), and legacy grapevines growing amongst
colonizing shrubs (c).

73
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

3.3.1.2 Sample analysis


The main set of samples were airdried for 48 hours and then sieved to procure the fine earth
fraction (< 2 mm) for analysis. Soil pH and electroconductivity were determined via multimeter
probe (1:2.5 deionized water) (MM 40+, Crison Instruments, Spain). All grinding and sample
preparation equipment was rinsed with acetone between each sample to reduce residual C and
isotopic signals. After pulverization and homogenization of the fine fraction with a mixer mill,
the samples were fumigated with HCl 32% to remove inorganic carbon contents and analyzed
for SOC, total nitrogen (N), and the ratio of stable isotopes 13C:12C (δ13C) and 15N:14N (δ15N)
via elemental analyzer (FlashEA 1112, Thermo Fisher Scientific, USA) coupled to an isotope
ratio mass spectrometer (DELTA V Advantage, Thermo Fisher Scientific, USA). To determine
total carbon (TC) and soil inorganic carbon (SIC), another complete set of samples was
analyzed without acid fumigation treatment. The natural abundance of stable isotopes was
expressed in delta (δ) notation with values in parts per mille (‰), relative to the international
Vienna Pee Dee Belemnite standard (IAEA, Austria) for 13C and atmospheric N for 15N (values
provided in Supplementary materials Table 6). The undisturbed (bulk density) samples were
oven dried at 105° C for 24 hours and weighed for fine soil and coarse rock fractions (Blake
and Hartge, 1986). Bulk density values for the fine soil fraction produced from equation (1)
were used to calculate SOC and N stock, based on an approximated rock fragments density of
2.6 g cm–3 (Don et al., 2007), using equation (2):

𝑀𝑠 − 𝑀𝑟𝑓 (1)
𝐵𝐷 =
𝑀𝑟𝑓
𝑉𝑠 − 𝜌
𝑟𝑓

(2)
𝑆𝑂𝐶𝑠𝑡 = 𝑆𝑂𝐶 × 𝐵𝐷 × 𝐷 × (1 − 𝑟𝑓𝑓)

where BD is the bulk density (g cm–3), Ms is the sample mass (g), Vs is the sample volume
(cm3), Mrf is the sample’s rock fragment mass (g), rf is the rock fragment density (g cm–3),
SOCst is the SOC (or N) stock (Mg ha–1), SOC is the concentration of SOC (or N) (%), D is
the soil depth investigated (cm), and rff is the gravimetric rock fragments fraction (vol. %/100).

3.3.2 Published data synthesis


Published chronosequence and paired plot studies undertaken in peninsular Spain investigating
the impacts of ALA on grassland, shrubland, and forest succession were compiled for analysis
following a literature search. While repeated measurements are the most ideal approach for
determining the effects of land use change over time, chronosequences and paired plots are

74
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

proven alternatives commonly employed (Breuer et al., 2006; Walker et al., 2010). Key terms
were searched using ISI Web of Science, Scopus, and Google Scholar with results limited to
English language studies conducted within peninsular Spain published since 1990. For an
individual study to be included, the time since abandonment (years) and the SOC or soil organic
matter (SOM) concentration (various units) within the topsoil (0–30 cm) of each
chronosequence stage and paired plot must have been provided. The following necessary
secondary criteria were either extracted from the studies themselves or determined through
other sources: mean annual precipitation (mm), mean annual temperature (° C), past crop type
(woody (e.g., Prunus dulcis, Olea europaea, Vitis vinifera) or annual (e.g., Hordeum vulgare,
Triticum aestivum)), and sampling site coordinates. Furthermore, each study included must
have additionally provided all the required data for a suitable agricultural control field (actively
cultivated, representing 0 years since abandonment). Studies were excluded if they used active
pastures or meadows as a control, if the past crop type was not explicitly stated, or if the data
points did not span at least more than 10 years post-abandonment. A total of 24 published
studies were identified under these criteria ranging from 1997–2020 (Supplementary materials
Table 8). With our field sites included (n = 12), the final dataset featured 113 examples of ALA
and 64 agricultural and natural control plots (Figure 12). SOC and SOM data were extracted
from tables, text, or by digitizing graphs (GetData Graph Digitizer, v.2.26, Russia). When only
SOM was provided (4 studies), SOC values were calculated from the Van Bemmelen
conversion factor following (Guo and Gifford, 2002)). This is permissible since only the
relative change in SOC over time is relevant and not the absolute SOC values. To standardize
the effect of time since abandonment on SOC content between chronosequences (La Mantia et
al., 2013), all SOC data points were plotted in reference to their paired agricultural control SOC
values according to equation (4):

𝑆𝑂𝐶𝑎𝑏 − 𝑆𝑂𝐶𝑎𝑔 (4)


𝛥𝑆𝑂𝐶 = × 100
𝑆𝑂𝐶𝑎𝑔

where ΔSOC is the relative change in SOC concentration (%), SOCab is SOC concentration of
the chronosequence stage investigated (%), and SOCag is the SOC concentration of the
associated agricultural control (%).

75
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

Figure 12. Locations of the published chronosequences and paired plots (circles) and our
chronosequences (triangles), representing 351 individual field plots in peninsular Spain.

3.3.3 Statistical analysis


Homoscedasticity of the data was verified with Levene's test and normality with the Shapiro–
Wilk test. Pearson correlation coefficients were calculated to identify relationships between the
soil variables. The effect of abandonment in the entire soil profile and per depth was assessed
for each of the chronosequences through one and two-way analysis of variance (ANOVA).
Non-normality in Torrelavit N stock data required a non-parametric Kruskal–Wallis test. When
statistical differences among means were apparent, a post-hoc multiple comparison test was
performed (Tukey-HSD). Data from all Spanish chronosequences were modelled (generalized
linear model with Akaike information criterion (AICc)) and fit to linear regressions to
determine the influence and interactions of time since abandonment, past crop type, MAT, and
MAP on SOC accumulation. Significance was assumed at p < .05 and all calculations and
visualizations were performed using R statistical software (R Core Team, 2020) and Grapher
(Golden Software, v.15, USA).

76
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

3.4 Results
3.4.1 SOC and N concentrations by successional stage and depth
Soil C concentration in the entire soil profile (0–30 cm) increased significantly from croplands
to either mid-stage or late-stage forests in all three chronosequences (Table 5). The mid-stage
forests, representing the most advanced stages of forest development post-abandonment in our
chronosequences, contained 358% (p = .005) and 148% (p = .004) higher concentrations of
SOC compared to the cropland control sites for Font-rubí and Cànoves, respectively, while
late-stage forests (the natural controls) contained 386% (p =.003) more SOC for Font-rubí and
120% (p = .021) for Torrelavit. ALA across all three chronosequences resulted in an average
SOC increase of 157% from croplands (1.10% SOC) to the mid-stage forests (2.82% SOC).

Soil N concentration also demonstrated an increasing trend over time following abandonment,
although not in all cases. The mid-stage forest for Font-rubí contained a 157% (p = .046) higher
N concentration compared to the cropland control site, while the same stage for Cànoves did
not differ significantly from the cropland control but contained 79% (p = .017) more N than
the early-stage forest. The stages of Torrelavit did not show any significant differences amongst
each other. On average, soil N concentrations increased by 51% from croplands (0.11% N) to
mid-stage forests (0.16% N).

77
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

Table 5. Mean soil chemical characteristics of the field sites (0–30 cm). Values in parentheses
represent standard error (±). Letters in columns indicate significant differences between stages
within each chronosequence (p < 0.05). Values per depth increment provided in Table 7.

Chronosequence Stage SOC (%) SIC (%) TC (%) N (%) EC (µS cm–1) pH

0.75 6.57 7.32 0.07


Cropland 165 8.14
(0.14)a (0.58) (0.44) (0.01)a

1.83 5.13 6.95 0.12


Early-stage 202 7.96
(0.21)ab (1.05) (0.86) (0.01)ab
Font-rubí
3.42 6.21 9.63 0.18
Mid-stage 250 7.79
(0.44)bc (1.39) (1.26) (0.02)b

3.63 5.86 9.80 0.19


Late-stage 249 7.7
(0.56)c (1.33) (0.53) (0.04)b

1.63 7.91 9.59


Cropland 0.12 (0.00) 227 8.03
(0.16)a (0.66) (0.70)

1.67 5.34 7.02


Early-stage 0.14 (0.01) 225 7.9
(0.17)a (0.86) (0.70)
Torrelavit
2.47 5.68 8.14
Mid-stage 0.13 (0.04) 253 7.74
(0.25)ab (0.36) (0.55)

3.58 6.29 9.87


Late-stage 0.18 (0.03) 273 7.64
(0.62)b (0.82) (0.99)

1.04 1.41 2.45 0.13


Cropland 190 7.62
(0.10)a (0.31) (0.39) (0.01)ab

1.36 3.26 4.62 0.09


Early-stage 157 7.72
(0.15)a (1.22) (1.26) (0.01)a
Cànoves
2.58 1.78 4.36 0.17
Mid-stage 187 7.65
(0.33)b (0.64) (0.97) (0.02)b

1.73 1.25 2.98 0.11


Late-stage 149 7.99
(0.20)ab (0.26) (0.19) (0.00)ab

Within the soil profile, the rates of decrease of SOC and N concentrations with increasing
sampling depth increased following ALA in all three chronosequences (Figure 13). Mid-stage
forests featured the highest average decrease for both SOC and N at 0.68 and 0.05 g kg–1 cm–1,
respectively. Conversely, cropland fields showed the smallest decrease with depth for SOC and

78
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

N at 0.17 and 0.02 g kg–1 cm–1, respectively. Differences in N concentrations among the stages
of succession in the surface soil (0–10 cm) rapidly converge down the soil profile, displaying
similar values in the 20–30 cm depth regardless of time since abandonment. This trend is also
noticeable to a lesser degree for SOC. The magnitude of both SOC and N changes as croplands
transitioned to mid-stage forests were highest at the soil surface, increasing by 188% and 99%
respectively. The lowest increases occurred in the 20–30 cm depth (125% for SOC and 15%
for N). Of all the soil variables among the three chronosequences, only SOC concentration in
Cànoves exhibited a significant interaction effect between stage and depth (p = .024).

Figure 13. Effect of soil depth (0–30 cm): (a) average change in SOC (g kg–1) by stage for all
three chronosequences; (b) average change in N (g kg–1) by stage for all three
chronosequences; (c) rate of decrease of SOC (g kg–1 cm–1) down the soil profile by stage for
each chronosequence; (d) rate of decrease of N (g kg–1 cm–1) down the soil profile by stage for
each chronosequence.

79
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

3.4.2 SOC and N stocks by successional stage and depth


Similar to SOC concentration, calculated SOC stocks in the entire soil profile increased from
cropland to mid-stage and late-stage forests in all three chronosequences (Figure 14). SOC
stock increased from cropland (23.7 Mg ha–1) to mid-stage forests by an average of 40.8 Mg
ha–1 (+172%) as a result of ALA, with significant increases for Font-rubí (+51.6 Mg ha–1, p =
.039) and Cànoves (+34.9 Mg ha–1, p = .013). Late-stage forests, representing the natural
control, featured an average SOC and soil N stock of 64.3 Mg ha–1 and 3.5 Mg ha–1,
respectively.

80
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

Figure 14. Change in SOC and N stock (Mg ha–1) for each chronosequence: (a, b) Font-rubí;
(c, d) Torrelavit; (e, f) Cànoves. Error bars represent standard error (±). Lowercase letters
above bars represent significant differences between depths within each stage, while uppercase
letters indicate statistical differences between stages within each chronosequence (p < 0.05).
Within the soil profile, SOC and N stock decreased with increasing sampling depth. Soil depth
played a significant role in Font-rubí on SOC stock only in the late-stage forest, and in both the
mid-stage and late-stage forests for N stock. Cànoves featured significant SOC and N
differences within the soil profiles of each successional stage, in contrast to Torrelavit. The
highest SOC and N stock values across all stages of all three chronosequences were observed

81
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

closest to the soil surface (0–10 cm), while the lowest values were virtually all found at the
lowest sampling depth (20–30 cm).

3.4.3 Soil C:N ratios by successional stage


Font-rubí and Cànoves showed an increasing soil C:N ratio from cropland to mid-stage forests
following abandonment (Figure 15). Cànoves exhibited significant increases in comparison to
croplands for each successional stage, with values of 14.2 (p = .001) for early-stage forests,
15.6 (p = .0004) for mid-stage forests, and 16.1 (p = .0002) for late-stage forests. Forest
regrowth had no significant effect on the C:N ratio across all chronosequence stages of
Torrelavit. The average C:N ratio of the three chronosequences increased from 10.5 (croplands)
to 18.4 (mid-stage). Average late-stage forests had a C:N ratio of 19.6.

Figure 15. C:N ratio of different chronosequence stages. Letters indicate significant
differences between stages within the same chronosequence (p < 0.05). Values per depth
increment provided in Supplementary materials Table 7.

3.4.4 Synthesis of chronosequences in Spain


Data from published chronosequences and paired plots, incorporated with the field data
presented above, indicates that ALA in Spain has had a net positive impact on promoting SCS.
The average rate of SOC accumulation following ALA is +2.3% yr–1 (R2 = 0.14, p < .0001)
(Figure 16.a) relative to cropland control fields, requiring nearly four and a half decades of

82
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

ecological succession before the first doubling of SOC concentration can be observed. The
synthesis featured 113 examples of ALA (after excluding cropland and natural controls) with
nearly 80% indicating a positive change. The average age across all sites is 24 years post-
abandonment and the average change in SOC is positive, at +69%. However, even after several
decades negative values have been reported due to various factors preventing a universally
positive trend over time for all categories of sites.

Figure 16. Relative change in SOC concentration (%) with time since abandonment (yr): (a)
all Spanish chronosequence sites; (b) Past crop type with linear regressions for sites previously
used for woody and annual crop production; (c) MAP (mm) with linear regressions for < 450
mm, 450 to < 1000 mm, and ≥ 1000 mm sites; (d) MAT (° C) with linear regressions for sites
within and outside a temperature range of 13–17 ° C. Shaded areas represent 95% confidence
interval. Horizontal line labelled “Natural control” represents average relative difference in
SOC concentration between paired natural and agricultural control sites (i.e., horizontal line
at 0).

83
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

Previous crop type is an important factor in SOC accumulation post-abandonment, with annual
increases of 4% for abandoned perennial woody croplands (R2 = 0.33, p < .0001, n = 61)
compared to annual cereal croplands (not significant, n = 48) (Figure 16.b). Less than 10% of
abandoned woody cropland sites demonstrated a negative change in SOC, compared to 40% of
annual cropland sites (Supplementary materials Figure 17). After three decades post-
abandonment, all woody cropland sites reported a gain in SOC. Similarly, sites below 1000
mm MAP displayed a positive trend in SOC accumulation over time, while sites at or above
this threshold exhibited no relationship (Figure 16.c). After two decades post-abandonment, all
observed losses of SOC were found in drier sites below 450 mm MAP (n = 48) and in more
humid sites ≥ 1000 mm MAP (n = 15). Within this precipitation range (n = 50), over 90% of
sites reported a positive change (reaching 100% after two decades) with an annual
accumulation rate of +3% (R2 = 0.36, p < .0001). Sites between 13 and 17 ° C (n = 64) also
demonstrated a greater SOC accumulation rate of +3% yr–1 (R2 = 0.27, p < .0001) compared to
sites above or below this range (not significant, n = 49) (Figure 16.d). Sites with observed SOC
losses were mainly outside this range (Supplementary materials Figure 17). The most important
variables for predicting SOC accumulation (model-averaged importance > 0.8) were time since
abandonment, past crop type, MAT and the variable interactions of past crop type with both
time and MAT (adj. R2 = 0.49, F7,105 = 16.47, p < 0.0001). Accumulation post-abandonment
was significantly correlated (bivariate) with time since abandonment (r = 0.38, p < .0001) and
past crop type (r = 0.32, p < .001) but not significantly with MAT (r = 0.10) nor MAP (r = -
0.13).

3.5 Discussion
3.5.1 Post-agricultural SOC and N changes with succession and depth
All our abandoned field sites exhibited higher SOC concentrations and stock values than active
cropland sites, as expected and observed across diverse ecosystems (Chiti et al., 2018; Deng et
al., 2016a; Gabarrón-Galeote et al., 2015a; Hu et al., 2018; Spohn et al., 2016). Our mid-stage
forests displayed a SOC stock gain of 40.8 Mg ha–1 (+172%) compared to the cropland controls.
This is higher than the 120% increases observed in the central Apennine range of Italy (+66.5
Mg ha–1) (Chiti et al., 2018) and in southwestern China under notably higher precipitation (>
1500 mm) (+50.6 Mg ha–1) (Yang et al., 2016). Differences in study parameters considered,
such as time since abandonment, climatic conditions, and restoration methods help explain the
range of SOC changes reported. With afforestation for example, SOC increases of 116±54%
have been reported in the temperate zone (Poeplau et al., 2011) and 190% across arid and semi-

84
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

arid regions (on abandoned and barren lands) (Liu et al., 2018). Intensive practices employed
in olive groves and vineyards across the Mediterranean are generally not conducive for SCS.
The cessation of crop residue removal (e.g., branches taken for firewood) and periodic soil
disturbances (e.g., tillage) can promote SOC accumulation (Debasish-Saha et al., 2014;
Laganière et al., 2010; Romero-Díaz et al., 2017). Revegetation reduces soil temperature, water
evaporation, and erosion while increasing the quantity and quality of organic matter inputs to
compensate losses from decomposition (Serpa et al., 2015). Post-abandonment increases in
fine root biomass also supports new SOC rhizodeposition (Novara et al., 2014), while fresh
plant residues promote macro-aggregate formation and therefore SOC stabilization and
accumulation (Nadal-Romero et al., 2016).

Although all three of our chronosequences showed positive SOC accumulation trends along
successional trajectories, significant changes from cropland values were not found until the
mid-stage forests. Similar lag times (∼30–50 years) between post-agricultural revegetation and
significant SOC and N increases have been identified in subtropical soils of southwestern China
(Hu et al., 2018; Yang et al., 2016) and following afforestation in a global perspective (Li et
al., 2012). The mechanisms that regulate SOC and N accumulation are different (McLauchlan,
2006), but N availability is considered a key factor in both the process of secondary succession
and the potential for simultaneous long-term SCS in soils post-abandonment (Johannes M H
Knops and Tilman, 2000; Li et al., 2012; Luo et al., 2004). The lack of any significant
differences in soil N stock between our chronosequence stages and the clear accumulation of
SOC throughout the successional process makes the presence of a strong N limiting effect
unlikely (Hu et al., 2018; Wen et al., 2016). Immediate above- and belowground plant biomass
production following abandonment may be permitted in the early stages of succession with
available N from remnant agricultural fertilizers (Spohn et al., 2016). The increasing soil C:N
ratio with each progressive successional stage has also been observed in other ALA
chronosequences studies (Deng et al., 2013; Li et al., 2012; Spohn et al., 2016). Because both
SOC and N increased from cropland to mid-stage forests, the positive correlation between C:N
and SOC accumulation can be explained by higher rates of C input than N during forest
regrowth (Deng et al., 2016a; Li et al., 2012).

Similar to other long-term ALA chronosequence studies (Zhao et al., 2015), SOC accumulation
also decreased with each successive stage in Font-rubí and Torrelavit. This might indicate a
SOC saturation effect influencing the sequestration capacity of mid- and late-stage forests
(Stewart et al., 2007). Conversely, it may also be a result of a proportionately higher

85
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

contribution of new SOC in early stages of succession (e.g., from new vegetation on soils
previously barren), while later stages exhibit a more balanced SOC budget resulting in a lower
SCS rate (i.e., steady inputs contribute proportionately less over time as SOC accumulates).
Soil depth exhibited a noticeable effect on SOC and N for all stages of succession of each
chronosequence. Converging soil N concentrations by the 20–30 cm depth among all stages of
succession indicates minimal depth penetration over time. As reported in other studies, the
highest concentrations of SOC and N in our sites were all in the 0–10 cm depth, where nutrient
accumulation throughout the successional process was highest (Hu et al., 2018; La Mantia et
al., 2013; Nadal-Romero et al., 2016). While cropland soils displayed the most homogenous
profiles due to tillage, SOC and N decreased with depth faster in mid-stage forests than late-
stage forests (natural controls presumably never tilled), indicating greater rates of surface
accumulation than depth saturation in regenerating post-agricultural soils. An opposite legacy
of tillage (i.e., distinct homogeneity) is also possible when excluding the effect of time since
abandonment (Sulman et al., 2020). Nutrient modelling efforts over regions containing both
previously tilled and never tilled revegetated soils should account for their potential differences
in depth distributions (e.g., custom pedotransfer functions) (Fernández-Ugalde and Tóth,
2017).

3.5.2 Post-agricultural SOC changes in Spain


Our study reveals that ALA has an overall net positive impact on promoting SCS in peninsular
Spain. The average SOC concentration accumulation rate of +2.3% y–1 varies depending on
site-specific conditions, as shown by negative rates reported even after several decades of
secondary succession. Nearly four and a half decades would be required before the first
doubling of SOC with a much longer period needed before reaching pre-agricultural levels.
This supports the view that SCS post-abandonment in Mediterranean environments can be a
slow, long-term process (Chiti et al., 2018; Lesschen et al., 2008; Nadal-Romero et al., 2016;
Segura et al., 2020). Spanish SOC stocks likely reach equilibrium well after the default 20-year
transition period assumed in carbon stock calculations following land use and land cover
changes (IPCC, 2006; Segura et al., 2020). In comparison, Paul et al., (2002) calculated a
relative SOC accumulation rate of 1.9% y–1 in topsoils following active restoration (i.e.,
afforestation) on former croplands globally, while Shi and Han, (2014)) calculated SOC
accumulation rates across China between 1.2–8.8% yr–1 during passive restoration (i.e., natural
succession), 10.7% yr–1 during natural grassland succession, and 0.8–4.7% yr–1 during
afforestation.

86
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

Mean annual precipitation at or above a threshold of approximately 1000 mm limited the


overall positive effect of time since abandonment on SOC concentration. High precipitation
inducing drop-offs in SCS have also been reported across Italy: Alberti et al., (2011) found
SOC accumulation below 900 mm MAP and losses above, while La Mantia et al., (2013))
found SOC gains at 650 mm and losses at 1100 mm during pasture-forest transitions. At or
above 1000 mm MAP in Spain, SOC losses have been reported in mountainous abandoned
agricultural lands in the north (Navas et al., 2012), and no gains after two decades in the south
(Gabarrón-Galeote et al., 2015b). High precipitation can result in N leaching and decreases in
aggregate protected SOC alongside increases in less protected particulate organic matter
fractions (Alberti et al., 2011; Guidi et al., 2014). Although long-term positive changes in SOC
post-abandonment at around 1000 mm MAP have also been observed in Mediterranean
environments (Chiti et al., 2018), precipitation and SOC accumulation during woody secondary
succession generally correlate negatively at the global scale (Jackson et al., 2002). In the
tropics, post-agricultural forests with a MAP below 1000 mm accumulate SOC faster than
forests with between 1000–2500 mm MAP, while no change can be expected in sites with
above 2500 mm MAP (Silver et al., 2001).

Accumulation of SOC on abandoned pastures and grazing lands transitioning to grasslands also
correlates positively with temperature and negatively with precipitation (Kämpf et al., 2016;
La Mantia et al., 2013; Pellis et al., 2019). The highest relative SCS rates have been in fact
observed within semi-arid climates (Kämpf et al., 2016; Liu et al., 2018). However,
precipitation levels closer to the lower limit of semi-arid conditions (i.e., < 450 mm MAP) can
also severely limit net primary productivity and therefore SOC accumulation through reduced
organic matter inputs (Figure 16.c) (Bonet, 2004; Gabarrón-Galeote et al., 2015b; Robledano-
Aymerich et al., 2014). Our results indicate a Goldilocks climate window of ~450–900 mm
MAP and ~13–17 ° C MAT for SCS during secondary succession post-abandonment in
Mediterranean environments. Effectively all of the sites in our synthesis within this window
showed net positive SOC changes, with increases of 100% and greater expected after three
decades. Our results also support the view that temperature plays a more dominant role in post-
abandonment SCS than previously thought (La Mantia et al., 2013; Pellis et al., 2019; Poeplau
et al., 2011), with precipitation playing more of an indirect role through its influence on other
factors (i.e., vegetation dynamics) (Wang et al., 2020).

Soil C stocks in Spain are also a function of past and present land management and land cover,
especially at local and regional scales (Hontoria et al., 1999; Muñoz-Rojas et al., 2015).

87
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

(Willaarts et al., 2016)) found a greater effect of land classification than MAT and MAP in the
south of Spain, with more SOC stock in croplands than in forests and shrublands because of
the flatter and deeper soil profiles of lands typically allocated for cultivation. During secondary
succession after ALA, our results indicate that the dominant historical crop type plays a
significant role in the rate of SOC accumulation. The high SCS capacity of abandoned woody
croplands has been observed throughout the Mediterranean region and elsewhere (Atallah et
al., 2015; Badalamenti et al., 2019; Romero-Díaz et al., 2017; Spohn et al., 2016). Soils used
for grain cultivation in semiarid regions of Spain may respond better to active restoration
compared to natural succession (Cuesta et al., 2012). Twenty-two years after abandoning cereal
fields in Andalusia, SOC gains were higher in afforested soils compared to naturally
regenerating soils (Segura et al., 2020). Our analysis featured very few active restoration sites
which may have limited the amount of positive values for semi-arid annual croplands (Garcia-
Franco et al., 2014); although this is not certain (Ruiz-Navarro et al., 2009).

Differences in the initial stock between woody and annual croplands at the time of
abandonment is likely the main reason for their different SOC accumulation rates. Vineyards
and orchards suffer from reduced SOM inputs (e.g., pruned branch removal) and are typically
allocated on marginal lands with lower quality soils (i.e., sloping, shallow, and stony) where
they grow better than cereal crops (García et al., 2007; Jebari et al., 2018; Pardini et al., 2003).
Cereal croplands may also receive SOM friendly management practices (e.g., regular manure
inputs, periodic stubble grazing and seed fallowing) which may entail SOC losses in the first
few years after their cessation (Navas et al., 2012; Ruecker et al., 1998). Recent estimations
indicate that topsoils of annual croplands have on average 6–7 Mg ha–1 more SOC than soils of
woody croplands in Spain (Calvo de Anta et al., 2020; Rodríguez Martín et al., 2016).
(Rodríguez-Murillo, 2001)) reported even greater differences between Spanish olive/vineyard
soils (40–43 Mg ha–1) and soils of irrigated/dryland croplands (51–58 Mg ha–1). The average
SOC concentrations for the woody croplands control sites in our synthesis were 10% lower
than the annual croplands. Fields with lower initial SOC are presumably farther from reaching
saturation and therefore have a higher carbon sink capacity during succession (Stewart et al.,
2007). Initial SOC stock and SCS potential correlate negatively during post-agricultural
succession globally (Deng et al., 2016b; Kämpf et al., 2016). In the Mediterranean and the
temperate zone, the sequestration potential of agricultural soils with > 50 Mg C ha–1 is limited
(Novara et al., 2017) or non-existent (Atallah et al., 2015; Kämpf et al., 2016).

88
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

3.5.3 Managing post-agricultural SOC accumulation under a changing


climate
Protecting and replenishing SOC stocks represents one quarter of the mitigation potential of all
land-based climate solutions (Bossio et al., 2020). Abandoned agricultural lands can be
managed in several ways that promote SCS while achieving other socioeconomic and
ecological goals (CHAPTER II; García-Ruiz et al., 2020; Yang et al., 2020). However,
Mediterranean precipitation and temperature regime changes will influence SOC accumulation
rates during secondary succession. In the south of Spain, where climate is recognized as one of
the most important drivers of ALA (Alonso-Sarría et al., 2016), drylands are expected to
expand (Gao and Giorgi, 2008) and precipitation to decline (-15%) (García-Ruiz et al., 2011).
Drier and hotter agricultural regions, such as in the southeast, face three future challenges:
greater losses of existing SOC stock (Jebari et al., 2018), greater risk of abandonment (Castillo
et al., 2020), and lower rates of SOC accumulation post-abandonment according to our results.
At the same time, predicted increases in precipitation intensity will exacerbate soil-plant water
stress (Rocha et al., 2020). This will impact abandoned land regeneration in addition to
irrigation agriculture which will already require more water following future infrastructure
modernization (Eekhout et al., 2018; Fader et al., 2016). Although abandonment typically
occurs on unproductive or difficult to access plots (i.e., marginal lands) (Rey Benayas et al.,
2007), a large percentage of abandoned land is highly suitable for forest growth and SOC
accumulation under drought conditions relative to undisturbed soils. This is due to their
relatively deep profiles with high available water capacities previously selected and favored
for tillage, and the possibility of legacy fertilizers buffering nutrient deficiencies during periods
of water stress (Willaarts et al., 2016). For example, secondary forests appearing in the second
half of the 20th century throughout Spain have been found to have higher growth rates than pre-
existing forests in drier regions (Vilà-Cabrera et al., 2017). However, higher evapotranspiration
resulting from the development of new forests on abandoned lands will further reduce surface
and sub-surface water resources.

The spontaneous regeneration of plant biomass and SOC following ALA implies multiple
climate change mitigation co-benefits in addition to removing atmospheric CO2 (Serpa et al.,
2015). Rural development strategies that intend to leverage ALA need to consider the high
variability of SOC responses and any potential risks that can offset intended benefits. As in
Italy (Novara et al., 2017), much of central Spain has experienced policy-driven and financially
incentivized abandonment of degraded cereal fields (Boellstorff and Benito, 2005). While the

89
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

intention was to improve soil conditions as seen in other parts of Europe, it has in some cases
increased erosion and therefore SOC losses, demonstrating the complexity and variability of
ALA impacts on Mediterranean soils (Rodrigo-Comino et al., 2018). Widespread unmanaged
forest regeneration in Mediterranean environments can also raise the risk of wildfires due to
increased plant homogeneity, biomass fuel, and forest connectivity (Viedma et al., 2006).
Planned or climate-induced crop conversions and land use changes, such as cereal production
in drier regions converted to bioenergy crops or left to regenerate into shrublands (Serpa et al.,
2015), presents additional opportunities for SCS that require further research efforts. The
ecological legacy of ALA in Spain and its potential for promoting land degradation neutrality
and climate change mitigation should be considered in rural development planning and policy-
making (van Leeuwen et al., 2019).

3.6 Conclusions
Agricultural land abandonment has produced divergent increases in SOC concentrations in
peninsular Spain. Chronosequence field studies indicate an average SOC accumulation rate of
+2.3% yr–1 post-abandonment. It is a highly variable process, depending on multiple
environmental and land management factors. The highest rates of SOC accumulation post-
abandonment can be expected on lands previously used for woody crop production featuring
~13–17 ° C MAT and ~450–900 mm MAP, with the lowest rates expected on lands previously
used for annual crop production outside this climatic window. Our secondary forest field sites
accrued 40.8 Mg C ha–1 (+172%) following abandonment but displayed greater SOC and N
depth heterogeneity than natural forests, demonstrating the long-lasting impact of agriculture.
By altering the SOC accumulation rates of existing secondary forests and influencing the
locations and crop types of future ALA, precipitation and temperature changes in the
Mediterranean region will determine the SCS potential and ecological value of abandoned
agricultural lands. Regional climate change mitigation policies in Mediterranean and semi-arid
environments can consider ALA as a low-cost but long-term option best incorporated in tandem
with other multipurpose sustainable land management strategies.

90
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

3.7 References
1. Alberti, G., Leronni, V., Piazzi, M., Petrella, F., Mairota, P., Peressotti, A., Piussi, P., Valentini, R., Gristina,
L., Mantia, T.L., Novara, A., Rühl, J., 2011. Impact of woody encroachment on soil organic carbon and
nitrogen in abandoned agricultural lands along a rainfall gradient in Italy. Reg. Environ. Chang.
https://doi.org/10.1007/s10113-011-0229-6

2. Alonso-Sarría, F., Martínez-Hernández, C., Romero-Díaz, A., Cánovas-García, F., Gomariz-Castillo, F.,
2016. Main Environmental Features Leading to Recent Land Abandonment in Murcia Region (Southeast
Spain). L. Degrad. Dev. 27, 654–670. https://doi.org/10.1002/ldr.2447

3. Arnaez, J., Lasanta, T., Errea, M.P., Ortigosa, L., 2011. Land abandonment, landscape evolution, and soil
erosion in a Spanish Mediterranean mountain region: The case of Camero Viejo. L. Degrad. Dev.
https://doi.org/10.1002/ldr.1032

4. Atallah, T., Sitt, K., El Asmar, E., Bitar, S., Ibrahim, L., Khatib, M.N., Darwish, T., 2015. Effect of
abandonment of olive orchards on soil organic carbon sequestration in Mediterranean Lebanon. Soil Res. 53,
745–752. https://doi.org/10.1071/SR14170

5. Badalamenti, E., Battipaglia, G., Gristina, L., Novara, A., Ruhl, J., Sala, G., Sapienza, L., Valentini, R., La
Mantia, T., 2019. Carbon stock increases up to old growth forest along a secondary succession in
Mediterranean island ecosystems. PLoS One 14. https://doi.org/10.1371/journal.pone.0220194

6. Baśnou, C., Álvarez, E., Bagaria, G., Guardiola, M., Isern, R., Vicente, P., Pino, J., 2013. Spatial patterns of
land use changes across a mediterranean metropolitan landscape: Implications for biodiversity management.
Environ. Manage. 52, 971–980. https://doi.org/10.1007/s00267-013-0150-5

7. Beck, H.E., Zimmermann, N.E., McVicar, T.R., Vergopolan, N., Berg, A., Wood, E.F., 2018. Present and
future köppen-geiger climate classification maps at 1-km resolution. Sci. Data 5.
https://doi.org/10.1038/sdata.2018.214

8. Beniston, J.W., DuPont, S.T., Glover, J.D., Lal, R., Dungait, J.A.J., 2014. Soil organic carbon dynamics 75
years after land-use change in perennial grassland and annual wheat agricultural systems. Biogeochemistry
120, 37–49. https://doi.org/10.1007/s10533-014-9980-3

9. Benjamin, K., Bouchard, A., Domon, G., 2007. Abandoned farmlands as components of rural landscapes: An
analysis of perceptions and representations. Landsc. Urban Plan. 83, 228–244.
https://doi.org/10.1016/j.landurbplan.2007.04.009

10. Blake, G.R., Hartge, K.H., 1986. Bulk Density, in: A. Klute et al. (Ed.), Methods of Soil Analysis: Part 1
Physical and Mineralogical Methods, 5.1. American Society of Agronomy ‐ Soil Science Society of America,
Madison, WI., pp. 363–375.

11. Boellstorff, D., Benito, G., 2005. Impacts of set-aside policy on the risk of soil erosion in central Spain. Agric.
Ecosyst. Environ. 107, 231–243. https://doi.org/10.1016/j.agee.2004.11.002

12. Bonet, A., 2004. Secondary succession of semi-arid Mediterranean old-fields in south-eastern Spain: insights
for conservation and restoration of degraded lands. J. Arid Environ. 56, 213–233.
https://doi.org/10.1016/S0140-1963(03)00048-X

13. Bossio, D.A., Cook-Patton, S.C., Ellis, P.W., Fargione, J., Sanderman, J., Smith, P., Wood, S., Zomer, R.J.,
von Unger, M., Emmer, I.M., Griscom, B.W., 2020. The role of soil carbon in natural climate solutions. Nat.
Sustain. https://doi.org/10.1038/s41893-020-0491-z

14. Breuer, L., Huisman, J.A., Keller, T., Frede, H.-G., 2006. Impact of a conversion from cropland to grassland
on C and N storage and related soil properties: Analysis of a 60-year chronosequence. Geoderma 133, 6–18.
https://doi.org/10.1016/J.GEODERMA.2006.03.033

15. Butzer, K.W., 2005. Environmental history in the Mediterranean world: Cross-disciplinary investigation of
cause-and-effect for degradation and soil erosion. J. Archaeol. Sci. 32, 1773–1800.
https://doi.org/10.1016/j.jas.2005.06.001

91
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

16. Calvo de Anta, R., Luís, E., Febrero-Bande, M., Galiñanes, J., Macías, F., Ortíz, R., Casás, F., 2020. Soil
organic carbon in peninsular Spain: Influence of environmental factors and spatial distribution. Geoderma
370. https://doi.org/10.1016/j.geoderma.2020.114365

17. Carlson, K.M., Gerber, J.S., Mueller, N.D., Herrero, M., MacDonald, G.K., Brauman, K.A., Havlik, P.,
O’Connell, C.S., Johnson, J.A., Saatchi, S., West, P.C., 2017. Greenhouse gas emissions intensity of global
croplands. Nat. Clim. Chang. 7, 63–68. https://doi.org/10.1038/nclimate3158

18. Castillo, C.P., Aliaga, E.C., Lavalle, C., Llario, J.C.M., 2020. An assessment and spatial modelling of
agricultural land abandonment in spain (2015-2030). Sustain. 12. https://doi.org/10.3390/su12020560

19. Chiti, T., Blasi, E., Pellis, G., Perugini, L., Chiriacò Maria, V., Valentini, R., 2018. Soil organic carbon pool’s
contribution to climate change mitigation on marginal land of a Mediterranean montane area in Italy. J.
Environ. Manage. https://doi.org/10.1016/j.jenvman.2018.04.093

20. Cohen, M., Varga, D., Vila, J., Barrassaud, E., 2011. A multi-scale and multi-disciplinary approach to monitor
landscape dynamics: A case study in the Catalan pre-Pyrenees (Spain). Geogr. J. 177, 79–91.
https://doi.org/10.1111/j.1475-4959.2010.00368.x

21. Cuesta, B., Rey Benayas, J.M., Gallardo, A., Villar-Salvador, P., González-Espinosa, M., 2012. Soil chemical
properties in abandoned Mediterranean cropland after succession and oak reforestation. Acta Oecologica 38,
58–65. https://doi.org/10.1016/j.actao.2011.09.004

22. De Baets, S., Van Oost, K., Baumann, K., Meersmans, J., Vanacker, V., Rumpel, C., 2012. Lignin signature
as a function of land abandonment and erosion in dry luvisols of SE Spain. CATENA 93, 78–86.
https://doi.org/10.1016/J.CATENA.2012.01.014

23. Debasish-Saha, Kukal, S.S., Bawa, S.S., 2014. Soil organic carbon stock and fractions in relation to land use
and soil depth in the degraded Shiwaliks hills of lower Himalayas. L. Degrad. Dev. 25, 407–416.
https://doi.org/10.1002/ldr.2151

24. Deng, L., Liu, G. bin, Shangguan, Z. ping, 2014. Land-use conversion and changing soil carbon stocks in
China’s “Grain-for-Green” Program: A synthesis. Glob. Chang. Biol. https://doi.org/10.1111/gcb.12508

25. Deng, L., Wang, K.-B., Chen, M.-L., Shangguan, Z.-P., Sweeney, S., 2013. Soil organic carbon storage
capacity positively related to forest succession on the Loess Plateau, China. CATENA 110, 1–7.
https://doi.org/10.1016/J.CATENA.2013.06.016

26. Deng, L., Wang, K., Tang, Z., Shangguan, Z., 2016a. Soil organic carbon dynamics following natural
vegetation restoration: Evidence from stable carbon isotopes (δ13C). Agric. Ecosyst. Environ. 221, 235–244.
https://doi.org/10.1016/J.AGEE.2016.01.048

27. Deng, L., Zhu, G. yu, Tang, Z. sheng, Shangguan, Z. ping, 2016b. Global patterns of the effects of land-use
changes on soil carbon stocks. Glob. Ecol. Conserv. https://doi.org/10.1016/j.gecco.2015.12.004

28. Don, A., Schumacher, J., Scherer-Lorenzen, M., Scholten, T., Schulze, E.D., 2007. Spatial and vertical
variation of soil carbon at two grassland sites - Implications for measuring soil carbon stocks. Geoderma 141,
272–282. https://doi.org/10.1016/j.geoderma.2007.06.003

29. Dunjó, G., Pardini, G., Gispert, M., 2003. Land use change effects on abandoned terraced soils in a
Mediterranean catchment, NE Spain. CATENA 52, 23–37. https://doi.org/10.1016/S0341-8162(02)00148-0

30. Eekhout, J.P.C., Hunink, J.E., Terink, W., De Vente, J., 2018. Why increased extreme precipitation under
climate change negatively affects water security. Hydrol. Earth Syst. Sci. 22, 5935–5946.
https://doi.org/10.5194/hess-22-5935-2018

31. Emran, M., Gispert, M., Pardini, G., 2012. Patterns of soil organic carbon, glomalin and structural stability
in abandoned Mediterranean terraced lands. Eur. J. Soil Sci. https://doi.org/10.1111/j.1365-
2389.2012.01493.x

92
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

32. Eugenio, M., Olano, J.M., Ferrandis, P., Martínez-Duro, E., Escudero, A., 2012. Population structure of two
dominant gypsophyte shrubs through a secondary plant succession. J. Arid Environ. 76, 30–35.
https://doi.org/10.1016/J.JARIDENV.2011.07.001

33. Faccioni, G., Sturaro, E., Ramanzin, M., Bernués, A., 2019. Socio-economic valuation of abandonment and
intensification of Alpine agroecosystems and associated ecosystem services. Land use policy 81, 453–462.
https://doi.org/10.1016/J.LANDUSEPOL.2018.10.044

34. Fader, M., Shi, S., Von Bloh, W., Bondeau, A., Cramer, W., 2016. Mediterranean irrigation under climate
change: More efficient irrigation needed to compensate for increases in irrigation water requirements. Hydrol.
Earth Syst. Sci. 20, 953–973. https://doi.org/10.5194/hess-20-953-2016

35. FAO, 2019. FAOSTAT Inputs/Land Use domain [WWW Document]. URL
http://www.fao.org/faostat/en/#data/RL (accessed 7.14.20).

36. Fernández-Ugalde, O., Tóth, G., 2017. Pedotransfer functions for predicting organic carbon in subsurface
horizons of European soils. Eur. J. Soil Sci. 68, 716–725. https://doi.org/10.1111/ejss.12464

37. Funes, I., Savé, R., Rovira, P., Molowny-Horas, R., Alcañiz, J.M., Ascaso, E., Herms, I., Herrero, C.,
Boixadera, J., Vayreda, J., 2019. Agricultural soil organic carbon stocks in the north-eastern Iberian
Peninsula: Drivers and spatial variability. Sci. Total Environ. 668, 283–294.
https://doi.org/10.1016/J.SCITOTENV.2019.02.317

38. Gabarrón-Galeote, M.A., Trigalet, S., van Wesemael, B., 2015a. Effect of land abandonment on soil organic
carbon fractions along a Mediterranean precipitation gradient. Geoderma 249–250, 69–78.
https://doi.org/10.1016/J.GEODERMA.2015.03.007

39. Gabarrón-Galeote, M.A., Trigalet, S., Wesemael, B. van, 2015b. Soil organic carbon evolution after land
abandonment along a precipitation gradient in southern Spain. Agric. Ecosyst. Environ. 199, 114–123.
https://doi.org/10.1016/J.AGEE.2014.08.027

40. Gao, X., Giorgi, F., 2008. Increased aridity in the Mediterranean region under greenhouse gas forcing
estimated from high resolution simulations with a regional climate model. Glob. Planet. Change 62, 195–209.
https://doi.org/10.1016/j.gloplacha.2008.02.002

41. Garcia-Franco, N., Wiesmeier, M., Goberna, M., Martínez-Mena, M., Albaladejo, J., 2014. Carbon dynamics
after afforestation of semiarid shrublands: Implications of site preparation techniques. For. Ecol. Manage.
319, 107–115. https://doi.org/10.1016/j.foreco.2014.01.043

42. García-Ruiz, J.M., Lasanta, T., Nadal-Romero, E., Lana-Renault, N., Álvarez-Farizo, B., 2020. Rewilding
and restoring cultural landscapes in Mediterranean mountains: Opportunities and challenges. Land use policy
99, 104850. https://doi.org/10.1016/j.landusepol.2020.104850

43. García-Ruiz, J.M., López-Moreno, I.I., Vicente-Serrano, S.M., Lasanta-Martínez, T., Beguería, S., 2011.
Mediterranean water resources in a global change scenario. Earth-Science Rev.
https://doi.org/10.1016/j.earscirev.2011.01.006

44. García, H., Tarrasón, D., Mayol, M., Male-Bascompte, N., Riba, M., 2007. Patterns of variability in soil
properties and vegetation cover following abandonment of olive groves in Catalonia (NE Spain). Acta
Oecologica 31, 316–324. https://doi.org/10.1016/J.ACTAO.2007.01.001

45. Guidi, C., Magid, J., Rodeghiero, M., Gianelle, D., Vesterdal, L., 2014. Effects of forest expansion on
mountain grassland: changes within soil organic carbon fractions. Plant Soil 385, 373–387.
https://doi.org/10.1007/s11104-014-2315-2

46. Guo, L.B., Gifford, R.M., 2002. Soil carbon stocks and land use change: A meta analysis. Glob. Chang. Biol.
https://doi.org/10.1046/j.1354-1013.2002.00486.x

47. Hontoria, C., Saa, A., Rodríguez-Murillo, J.C., 1999. Relationships Between Soil Organic Carbon and Site
Characteristics in Peninsular Spain. Soil Sci. Soc. Am. J. 63, 614–621.
https://doi.org/10.2136/sssaj1999.03615995006300030026x

93
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

48. Hoogmoed, M., Cunningham, S.C., Thomson, J.R., Baker, P.J., Beringer, J., Cavagnaro, T.R., 2012. Does
afforestation of pastures increase sequestration of soil carbon in Mediterranean climates? Agric. Ecosyst.
Environ. 159, 176–183. https://doi.org/10.1016/J.AGEE.2012.07.011

49. Hu, P., Liu, S., Ye, Y., Zhang, W., He, X., Su, Y., Wang, K., 2018. Soil carbon and nitrogen accumulation
following agricultural abandonment in a subtropical karst region. Appl. Soil Ecol.
https://doi.org/10.1016/j.apsoil.2018.09.003

50. IPCC, 2006. IPCC Guidelines for National Greenhouse Gas Inventories Volume 4: Agriculture, Forestry, and
Other Landuse. Hayama, Japan.

51. Jackson, R.B., Banner, J.L., Jobbágy, E.G., Pockman, W.T., Wall, D.H., 2002. Ecosystem carbon loss with
woody plant invasion of grasslands. Nature 418, 623–626. https://doi.org/10.1038/nature00910

52. Jebari, A., del Prado, A., Pardo, G., Rodríguez Martín, J.A., Álvaro-Fuentes, J., 2018. Modeling Regional
Effects of Climate Change on Soil Organic Carbon in Spain. J. Environ. Qual. 47, 644–653.
https://doi.org/10.2134/jeq2017.07.0294

53. Kämpf, I., Hölzel, N., Störrle, M., Broll, G., Kiehl, K., 2016. Potential of temperate agricultural soils for
carbon sequestration: A meta-analysis of land-use effects. Sci. Total Environ. 566–567, 428–435.
https://doi.org/10.1016/J.SCITOTENV.2016.05.067

54. Knops, J.M.H., Tilman, D., 2000. DYNAMICS OF SOIL NITROGEN AND CARBON ACCUMULATION
FOR 61 YEARS AFTER AGRICULTURAL ABANDONMENT, Ecology.

55. Kurganova, I., Lopes de Gerenyu, V., Six, J., Kuzyakov, Y., 2014. Carbon cost of collective farming collapse
in Russia. Glob. Chang. Biol. https://doi.org/10.1111/gcb.12379

56. La Mantia, T., Gristina, L., Rivaldo, E., Pasta, S., Novara, A., Rühl, J., 2013. The effects of post-pasture
woody plant colonization on soil and aboveground litter carbon and nitrogen along a bioclimatic transect.
IForest 6, 238–246. https://doi.org/10.3832/ifor0811-006

57. Laganière, J., Angers, D.A., Paré, D., 2010. Carbon accumulation in agricultural soils after afforestation: A
meta-analysis. Glob. Chang. Biol. 16, 439–453. https://doi.org/10.1111/j.1365-2486.2009.01930.x

58. Lal, R., 2013. Intensive Agriculture and the Soil Carbon Pool. J. Crop Improv. 27, 735–751.
https://doi.org/10.1080/15427528.2013.845053

59. Lasanta, T., Arnáez, J., Pascual, N., Ruiz-Flaño, P., Errea, M.P., Lana-Renault, N., 2017. Space–time process
and drivers of land abandonment in Europe. CATENA 149, 810–823.
https://doi.org/10.1016/J.CATENA.2016.02.024

60. Lesschen, J.P., Cammeraat, L.H., Kooijman, A.M., van Wesemael, B., 2008. Development of spatial
heterogeneity in vegetation and soil properties after land abandonment in a semi-arid ecosystem. J. Arid
Environ. 72, 2082–2092. https://doi.org/10.1016/J.JARIDENV.2008.06.006

61. Li, D., Niu, S., Luo, Y., 2012. Global patterns of the dynamics of soil carbon and nitrogen stocks following
afforestation: A meta-analysis. New Phytol. 195, 172–181. https://doi.org/10.1111/j.1469-
8137.2012.04150.x

62. Liu, X., Yang, T., Wang, Q., Huang, F., Li, L., 2018. Dynamics of soil carbon and nitrogen stocks after
afforestation in arid and semi-arid regions: A meta-analysis. Sci. Total Environ. 618, 1658–1664.
https://doi.org/10.1016/J.SCITOTENV.2017.10.009

63. Luo, Y., Su, B., Currie, W.S., Dukes, J.S., Finzi, A., Hartwig, U., Hungate, B., McMurtrie, R.E., Oren, R.,
Parton, W.J., Pataki, D.E., Shaw, R.M., Zak, D.R., Field, C.B., 2004. Progressive Nitrogen Limitation of
Ecosystem Responses to Rising Atmospheric Carbon Dioxide. Bioscience 54, 731–739.
https://doi.org/10.1641/0006-3568(2004)054[0731:PNLOER]2.0.CO;2

94
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

64. Martinez-Duro, E., Ferrandis, P., Escudero, A., Luzuriaga, A.L., Herranz, J.M., 2010. Secondary old-field
succession in an ecosystem with restrictive soils: Does time from abandonment matter? Appl. Veg. Sci.
https://doi.org/10.1111/j.1654-109X.2009.01064.x

65. McLauchlan, K., 2006. The nature and longevity of agricultural impacts on soil carbon and nutrients: A
review. Ecosystems. https://doi.org/10.1007/s10021-005-0135-1

66. Muñoz-Rojas, M., Jordán, A., Zavala, L.M., De la Rosa, D., Abd-Elmabod, S.K., Anaya-Romero, M., 2015.
Impact of Land Use and Land Cover Changes on Organic Carbon Stocks in Mediterranean Soils (1956-2007).
L. Degrad. Dev. 26, 168–179. https://doi.org/10.1002/ldr.2194

67. Nadal-Romero, E., Cammeraat, E., Pérez-Cardiel, E., Lasanta, T., 2016. How do soil organic carbon stocks
change after cropland abandonment in Mediterranean humid mountain areas? Sci. Total Environ. 566–567,
741–752. https://doi.org/10.1016/J.SCITOTENV.2016.05.031

68. Navas, A., Gaspar, L., Quijano, L., López-Vicente, M., Machín, J., 2012. Patterns of soil organic carbon and
nitrogen in relation to soil movement under different land uses in mountain fields (South Central Pyrenees).
Catena 94, 43–52. https://doi.org/10.1016/j.catena.2011.05.012

69. Novara, A., Gristina, L., La Mantia, T., Rühl, J., 2013. Carbon dynamics of soil organic matter in bulk soil
and aggregate fraction during secondary succession in a Mediterranean environment. Geoderma 193–194,
213–221. https://doi.org/10.1016/J.GEODERMA.2012.08.036

70. Novara, A., Gristina, L., Sala, G., Galati, A., Crescimanno, M., Cerdà, A., Badalamenti, E., La Mantia, T.,
2017. Agricultural land abandonment in Mediterranean environment provides ecosystem services via soil
carbon sequestration. Sci. Total Environ. https://doi.org/10.1016/j.scitotenv.2016.10.123

71. Novara, A., La Mantia, T., Rühl, J., Badalucco, L., Kuzyakov, Y., Gristina, L., Laudicina, V.A., 2014.
Dynamics of soil organic carbon pools after agricultural abandonment. Geoderma 235–236, 191–198.
https://doi.org/10.1016/j.geoderma.2014.07.015

72. Padilla, F.M., Vidal, B., Sánchez, J., Pugnaire, F.I., 2010. Land-use changes and carbon sequestration through
the twentieth century in a Mediterranean mountain ecosystem: Implications for land management. J. Environ.
Manage. 91, 2688–2695. https://doi.org/10.1016/J.JENVMAN.2010.07.031

73. Pardini, G., Gispert, M., Dunjó, G., 2003. Runoff erosion and nutrient depletion in five Mediterranean soils
of NE Spain under different land use. Sci. Total Environ. 309, 213–224. https://doi.org/10.1016/S0048-
9697(03)00007-X

74. Paul, K.I., Polglase, P.J., Nyakuengama, J.G., Khanna, P.K., 2002. Change in soil carbon following
afforestation. For. Ecol. Manage. 168, 241–257. https://doi.org/10.1016/S0378-1127(01)00740-X

75. Pausas, J.G., Millán, M.M., 2019. Greening and Browning in a Climate Change Hotspot: The Mediterranean
Basin. Bioscience. https://doi.org/10.1093/biosci/biy157

76. Pellis, G., Chiti, T., Rey, A., Curiel Yuste, J., Trotta, C., Papale, D., 2019. The ecosystem carbon sink
implications of mountain forest expansion into abandoned grazing land: The role of subsoil and climatic
factors. Sci. Total Environ. 672, 106–120. https://doi.org/10.1016/J.SCITOTENV.2019.03.329

77. Poeplau, C., Don, A., Vesterdal, L., Leifeld, J., Van Wesemael, B., Schumacher, J., Gensior, A., 2011.
Temporal dynamics of soil organic carbon after land-use change in the temperate zone - carbon response
functions as a model approach. Glob. Chang. Biol. https://doi.org/10.1111/j.1365-2486.2011.02408.x

78. Post, W.M., Kwon, K.C., 2000. Soil carbon sequestration and land‐use change: processes and potential. Glob.
Chang. Biol. 6. https://doi.org/10.1046/j.1365-2486.2000.00308.x

79. R Core Team, 2020. R: A language and environment for statistical computing. R Foundation for Statistical
Computing.

95
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

80. Rey Benayas, J.M., Martins, A., Nicolau, J.M., Schulz, J.J., 2007. Abandonment of agricultural land: an
overview of drivers and consequences. CAB Rev. Perspect. Agric. Vet. Sci. Nutr. Nat. Resour.
https://doi.org/10.1079/PAVSNNR20072057

81. Robledano-Aymerich, F., Romero-Díaz, A., Belmonte-Serrato, F., Zapata-Pérez, V.M., Martínez-Hernández,
C., Martínez-López, V., 2014. Ecogeomorphological consequences of land abandonment in semiarid
Mediterranean areas: Integrated assessment of physical evolution and biodiversity. Agric. Ecosyst. Environ.
197, 222–242. https://doi.org/10.1016/J.AGEE.2014.08.006

82. Rocha, J., Carvalho-Santos, C., Diogo, P., Beça, P., Keizer, J.J., Nunes, J.P., 2020. Impacts of climate change
on reservoir water availability, quality and irrigation needs in a water scarce Mediterranean region (southern
Portugal). Sci. Total Environ. 736. https://doi.org/10.1016/j.scitotenv.2020.139477

83. Rodrigo-Comino, J., Martínez-Hernández, C., Iserloh, T., Cerdà, A., 2018. Contrasted Impact of Land
Abandonment on Soil Erosion in Mediterranean Agriculture Fields. Pedosphere.
https://doi.org/10.1016/S1002-0160(17)60441-7

84. Rodríguez-Murillo, J.C., 2001. Organic carbon content under different types of land use and soil in peninsular
Spain, Biol Fertil Soils. Springer-Verlag.

85. Rodríguez Martín, J.A., Álvaro-Fuentes, J., Gonzalo, J., Gil, C., Ramos-Miras, J.J., Grau Corbí, J.M., Boluda,
R., 2016. Assessment of the soil organic carbon stock in Spain. Geoderma 264, 117–125.
https://doi.org/10.1016/J.GEODERMA.2015.10.010

86. Romero-Díaz, A., Ruiz-Sinoga, J.D., Robledano-Aymerich, F., Brevik, E.C., Cerdà, A., 2017. Ecosystem
responses to land abandonment in Western Mediterranean Mountains. Catena.
https://doi.org/10.1016/j.catena.2016.08.013

87. Ruecker, G., Schad, P., Alcubilla, M.M., Ferrer, C., 1998. Natural regeneration of degraded soils and site
changes on abandoned agricultural terraces in Mediterranean Spain. L. Degrad. Dev. 9, 179–188.

88. Ruiz-Navarro, A., Barberá, G.G., Navarro-Cano, J.A., Albaladejo, J., Castillo, V.M., 2009. Soil dynamics in
Pinus halepensis reforestation: Effect of microenvironments and previous land use. Geoderma 153, 353–361.
https://doi.org/10.1016/j.geoderma.2009.08.024

89. Ruskule, A., Nikodemus, O., Kasparinskis, R., Bell, S., Urtane, I., 2013. The perception of abandoned
farmland by local people and experts: Landscape value and perspectives on future land use. Landsc. Urban
Plan. 115, 49–61. https://doi.org/10.1016/J.LANDURBPLAN.2013.03.012

90. Segura, C., Navarro, F.B., Jiménez, M.N., Fernández-Ondoño, E., 2020. Implications of afforestation vs.
secondary succession for soil properties under a semiarid climate. Sci. Total Environ. 704.
https://doi.org/10.1016/j.scitotenv.2019.135393

91. Serpa, D., Nunes, J.P., Santos, J., Sampaio, E., Jacinto, R., Veiga, S., Lima, J.C., Moreira, M., Corte-Real, J.,
Keizer, J.J., Abrantes, N., 2015. Impacts of climate and land use changes on the hydrological and erosion
processes of two contrasting Mediterranean catchments. Sci. Total Environ. 538, 64–77.
https://doi.org/10.1016/j.scitotenv.2015.08.033

92. Shi, S., Han, P., 2014. Estimating the soil carbon sequestration potential of China’s Grain for Green Project.
Global Biogeochem. Cycles 28, 1279–1294. https://doi.org/10.1002/2014GB004924

93. Silver, W.L., Ostertag, R., Lugo, A.E., 2001. The Potential for Carbon Sequestration Through Reforestation
of Abandoned Tropical Agricultural and Pasture Lands.

94. Spohn, M., Novák, T.J., Incze, J., Giani, L., 2016. Dynamics of soil carbon, nitrogen, and phosphorus in
calcareous soils after land-use abandonment – A chronosequence study. Plant Soil.
https://doi.org/10.1007/s11104-015-2513-6

95. Stewart, C.E., Paustian, K., Conant, R.T., Plante, A.F., Six, J., 2007. Soil carbon saturation: Concept,
evidence and evaluation. Biogeochemistry 86, 19–31. https://doi.org/10.1007/s10533-007-9140-0

96
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

96. Stolbovoy, V., Montanarella, L., Filippi, A.J., Gallego, J., Grassi, G., 2007. Soil sampling protocol to certify
the changes of organic carbon stock in mineral soil of the European Union. Version 2. EUR 21576 EN/2.
Office for Official Publications of the European Communities, Luxembourg.

97. Sulman, B.N., Harden, J., He, Y., Treat, C., Koven, C., Mishra, U., O’Donnell, J.A., Nave, L.E., 2020. Land
Use and Land Cover Affect the Depth Distribution of Soil Carbon: Insights From a Large Database of Soil
Profiles. Front. Environ. Sci. 8. https://doi.org/10.3389/fenvs.2020.00146

98. Ustaoglu, E., Collier, M.J., 2018. Farmland abandonment in Europe: an overview of drivers, consequences,
and assessment of the sustainability implications. Environ. Rev. 26, 396–416. https://doi.org/10.1139/er-
2018-0001

99. van der Zanden, E.H., Carvalho-Ribeiro, S.M., Verburg, P.H., 2018. Abandonment landscapes: user attitudes,
alternative futures and land management in Castro Laboreiro, Portugal. Reg. Environ. Chang. 1–12.
https://doi.org/10.1007/s10113-018-1294-x

100. van Leeuwen, C.C.E., Cammeraat, E.L.H., de Vente, J., Boix-Fayos, C., 2019. The evolution of soil
conservation policies targeting land abandonment and soil erosion in Spain: A review. Land use policy 83,
174–186. https://doi.org/10.1016/J.LANDUSEPOL.2019.01.018

101. Viedma, O., Moreno, J.M., Rieiro, I., 2006. Interactions between land use/land cover change, forest fires and
landscape structure in Sierra de Gredos (Central Spain). Environ. Conserv. 33, 212–222.
https://doi.org/10.1017/S0376892906003122

102. Vilà-Cabrera, A., Espelta, J.M., Vayreda, J., Pino, J., 2017. “New Forests” from the Twentieth Century are a
Relevant Contribution for C Storage in the Iberian Peninsula. Ecosystems. https://doi.org/10.1007/s10021-
016-0019-6

103. Walker, L.R., Wardle, D.A., Bardgett, R.D., Clarkson, B.D., 2010. The use of chronosequences in studies of
ecological succession and soil development. J. Ecol. https://doi.org/10.1111/j.1365-2745.2010.01664.x

104. Wang, H., Yue, C., Mao, Q., Zhao, J., Ciais, P., Li, W., Yu, Q., Mu, X., 2020. Vegetation and species impacts
on soil organic carbon sequestration following ecological restoration over the Loess Plateau, China.
Geoderma 371. https://doi.org/10.1016/j.geoderma.2020.114389

105. Wen, L., Li, D., Yang, L., Luo, P., Chen, Hao, Xiao, K., Song, T., Zhang, W., He, X., Chen, Hongsong,
Wang, K., 2016. Rapid recuperation of soil nitrogen following agricultural abandonment in a karst area,
southwest China. Biogeochemistry 129, 341–354. https://doi.org/10.1007/s10533-016-0235-3

106. Wertebach, T.M., Hölzel, N., Kämpf, I., Yurtaev, A., Tupitsin, S., Kiehl, K., Kamp, J., Kleinebecker, T.,
2017. Soil carbon sequestration due to post-Soviet cropland abandonment: estimates from a large-scale soil
organic carbon field inventory. Glob. Chang. Biol. 23, 3729–3741. https://doi.org/10.1111/gcb.13650

107. Willaarts, B.A., Oyonarte, C., Muñoz-Rojas, M., Ibáñez, J.J., Aguilera, P.A., 2016. Environmental Factors
Controlling Soil Organic Carbon Stocks in Two Contrasting Mediterranean Climatic Areas of Southern
Spain. L. Degrad. Dev. https://doi.org/10.1002/ldr.2417

108. Yang, L., Luo, P., Wen, L., Li, D., 2016. Soil organic carbon accumulation during post-agricultural succession
in a karst area, southwest China. Sci. Rep. 6. https://doi.org/10.1038/srep37118

109. Yang, Y., Hobbie, S.E., Hernandez, R.R., Fargione, J., Grodsky, S.M., Tilman, D., Zhu, Y.-G., Luo, Y.,
Smith, T.M., Jungers, J.M., Yang, M., Chen, W.-Q., 2020. Restoring Abandoned Farmland to Mitigate
Climate Change on a Full Earth. One Earth 3, 176–186. https://doi.org/10.1016/j.oneear.2020.07.019

110. Zethof, J.H.T., Cammeraat, E.L.H., Nadal-Romero, E., 2019. The enhancing effect of afforestation over
secondary succession on soil quality under semiarid climate conditions. Sci. Total Environ. 652, 1090–1101.
https://doi.org/10.1016/J.SCITOTENV.2018.10.235

111. Zhao, Y.G., Liu, X.F., Wang, Z.L., Zhao, S.W., 2015. Soil organic carbon fractions and sequestration across
a 150-yr secondary forest chronosequence on the Loess Plateau, China. Catena 133, 303–308.
https://doi.org/10.1016/j.catena.2015.05.028

97
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

3.8 Supplementary materials


Table 6. Mean soil 13C and 15N stable isotope values of the field sites (0–30 cm). Values in
parentheses represent standard error (±). Letters in columns indicate significant differences
between stages within each chronosequence (p < 0.05).

Chronosequence Stage δ13C (‰) δ15N (‰)

Cropland -24.17 (1.39) 4.76 (0.34)

Early-stage -26.05 (0.12) 4.41 (0.94)


Font-rubí
Mid-stage -26.49 (1.13) 2.15 (1.30)

Late-stage -25.50 (1.03) 2.17 (0.14)

Cropland -20.45 (2.18) 5.25 (0.35)c

Early-stage -25.30 (0.06) 3.48 (0.19)b


Torrelavit
Mid-stage -23.01 (2.25) 1.05 (0.49)a

Late-stage -24.00 (0.97) 2.17 (0.05)ab

Cropland -25.67 (0.03)b 9.03 (0.35)c

Early-stage -26.25 (0.29)ab 0.15 (0.38)a


Cànoves
Mid-stage -26.34 (0.16)ab 1.72 (0.12)ab

Late-stage -26.55 (0.13)a 4.71 (1.43)b

98
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

Table 7. Mean soil chemical characteristics of the field sites by depth (n = 3).

0–10 cm 10–20 cm 20–30 cm

SOC SIC TC N C:N SOC SIC TC N C:N SOC SIC TC N C:N


Chronosequence Stage
(%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%)

Cropland 0.98 6.41 7.39 0.08 12.17 0.70 6.58 7.28 0.07 9.62 0.56 6.73 7.29 0.05 10.29

Early-stage 2.26 5.22 7.48 0.16 14.15 1.93 4.83 6.76 0.12 16.48 1.29 5.32 6.62 0.09 13.84
Font-rubí
Mid-stage 4.56 6.17 10.73 0.27 17.16 3.26 5.20 8.46 0.16 20.07 2.45 7.25 9.71 0.11 21.90

Late-stage 4.80 6.34 11.15 0.29 16.88 3.54 5.60 9.14 0.17 21.52 2.47 6.66 9.13 0.11 33.17

Cropland 1.75 8.45 10.20 0.13 14.07 1.68 7.98 9.52 0.12 18.85 1.39 6.83 9.05 0.12 18.26

Early-stage 2.02 5.51 7.53 0.17 12.28 1.51 5.07 6.58 0.13 11.65 1.49 5.45 6.94 0.13 11.20
Torrelavit
Mid-stage 3.30 6.37 9.67 0.22 15.78 2.19 5.32 7.51 0.10 25.40 1.91 5.34 7.25 0.07 20.15

Late-stage 4.14 7.19 11.33 0.23 18.06 3.74 5.83 9.57 0.18 21.60 2.85 5.85 8.70 0.14 19.49

Cropland 1.43 1.19 2.62 0.17 8.29 1.03 1.56 2.59 0.13 8.10 0.65 1.47 2.12 0.08 8.42

Early-stage 2.32 3.24 5.56 0.16 14.05 1.32 3.20 4.52 0.09 14.96 0.44 3.34 3.78 0.03 13.68
Cànoves
Mid-stage 4.11 3.05 7.16 0.28 15.28 2.11 1.07 3.18 0.13 15.75 1.50 1.24 2.74 0.10 14.72

Late-stage 2.82 2.32 5.14 0.20 14.21 1.49 0.59 2.08 0.09 17.40 0.88 0.86 1.73 005 16.56

99
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

Figure 17. Notched boxplots for past crop type, MAP, and MAT variables considered in the
synthesis of chronosequence data. Middle line represents median and whiskers represent upper
(90%) and lower (10%) percentiles.

100
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

Table 8. List of papers included in the Spanish synthesis.

1. Blanco-Moure, N., Gracia, R., Bielsa, A.C., López, M.V., 2016. Soil organic matter
fractions as affected by tillage and soil texture under semiarid Mediterranean
conditions. Soil Tillage Res. 155, 381–389.
https://doi.org/10.1016/J.STILL.2015.08.011

2. Bonet, A., 2004. Secondary succession of semi-arid Mediterranean old-fields in south-


eastern Spain: insights for conservation and restoration of degraded lands. J. Arid
Environ. 56, 213–233. https://doi.org/10.1016/S0140-1963(03)00048-X

3. Cammeraat, L.., Imeson, A.., 1999. The evolution and significance of soil–vegetation
patterns following land abandonment and fire in Spain. CATENA 37, 107–127.
https://doi.org/10.1016/S0341-8162(98)00072-1

4. De Baets, S., Meersmans, J., Vanacker, V., Quine, T.A., Van Oost, K., 2013. Spatial
variability and change in soil organic carbon stocks in response to recovery following
land abandonment and erosion in mountainous drylands. Soil Use Manag.
https://doi.org/10.1111/sum.12017

5. De Baets, S., Van Oost, K., Baumann, K., Meersmans, J., Vanacker, V., Rumpel, C.,
2012. Lignin signature as a function of land abandonment and erosion in dry luvisols
of SE Spain. CATENA 93, 78–86. https://doi.org/10.1016/J.CATENA.2012.01.014

6. Dunjó, G., Pardini, G., Gispert, M., 2003. Land use change effects on abandoned
terraced soils in a Mediterranean catchment, NE Spain. CATENA 52, 23–37.
https://doi.org/10.1016/S0341-8162(02)00148-0

7. Emran, M., Gispert, M., Pardini, G., 2012. Patterns of soil organic carbon, glomalin
and structural stability in abandoned Mediterranean terraced lands. Eur. J. Soil Sci.
https://doi.org/10.1111/j.1365-2389.2012.01493.x

8. Gabarrón-Galeote, M.A., Trigalet, S., van Wesemael, B., 2015. Effect of land
abandonment on soil organic carbon fractions along a Mediterranean precipitation
gradient. Geoderma 249–250, 69–78.
https://doi.org/10.1016/J.GEODERMA.2015.03.007

9. García, H., Tarrasón, D., Mayol, M., Male-Bascompte, N., Riba, M., 2007. Patterns of
variability in soil properties and vegetation cover following abandonment of olive
groves in Catalonia (NE Spain). Acta Oecologica 31, 316–324.
https://doi.org/10.1016/J.ACTAO.2007.01.001

10. Hontoria, C., Velásquez, R., Benito, M., Almorox, J., Moliner, A., 2009. Bradford-
reactive soil proteins and aggregate stability under abandoned versus tilled olive groves
in a semi-arid calcisol. Soil Biol. Biochem. 41, 1583–1585.
https://doi.org/10.1016/J.SOILBIO.2009.04.025

11. Lesschen, J.P., Cammeraat, L.H., Kooijman, A.M., van Wesemael, B., 2008.
Development of spatial heterogeneity in vegetation and soil properties after land

101
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

abandonment in a semi-arid ecosystem. J. Arid Environ. 72, 2082–2092.


https://doi.org/10.1016/J.JARIDENV.2008.06.006

12. Martinez-Mena, M., Lopez, J., Almagro, M., Boix-Fayos, C., Albaladejo, J., 2008.
Effect of water erosion and cultivation on the soil carbon stock in a semiarid area of
South-East Spain. Soil Tillage Res. 99, 119–129.
https://doi.org/10.1016/J.STILL.2008.01.009

13. Moreno, J.L., Torres, I.F., García, C., López-Mondéjar, R., Bastida, F., 2019. Land use
shapes the resistance of the soil microbial community and the C cycling response to
drought in a semi-arid area. Sci. Total Environ. 648, 1018–1030.
https://doi.org/10.1016/J.SCITOTENV.2018.08.214

14. Navas, A., Gaspar, L., Quijano, L., López-Vicente, M., Machín, J., 2012. Patterns of
soil organic carbon and nitrogen in relation to soil movement under different land uses
in mountain fields (South Central Pyrenees). Catena 94, 43–52.
https://doi.org/10.1016/j.catena.2011.05.012

15. Pardini, G., Gispert, M., Dunjó, G., 2003. Runoff erosion and nutrient depletion in five
Mediterranean soils of NE Spain under different land use. Sci. Total Environ. 309, 213–
224. https://doi.org/10.1016/S0048-9697(03)00007-X

16. Robledano-Aymerich, F., Romero-Díaz, A., Belmonte-Serrato, F., Zapata-Pérez, V.M.,


Martínez-Hernández, C., Martínez-López, V., 2014. Ecogeomorphological
consequences of land abandonment in semiarid Mediterranean areas: Integrated
assessment of physical evolution and biodiversity. Agric. Ecosyst. Environ. 197, 222–
242. https://doi.org/10.1016/J.AGEE.2014.08.006

17. Rodrigo-Comino, J., Martínez-Hernández, C., Iserloh, T., Cerdà, A., 2018. Contrasted
Impact of Land Abandonment on Soil Erosion in Mediterranean Agriculture Fields.
Pedosphere. https://doi.org/10.1016/S1002-0160(17)60441-7

18. Roldan, A., Garcia, C., Albaladejo, J., 1997. AM fungal abundance and activity in a
chronosequence of abandoned fields in a semiarid mediterranean site. Arid Soil Res.
Rehabil. https://doi.org/10.1128/IAI.01216-06

19. Romanyà, J., Cortina, J., Falloon, P., Coleman, K., Smith, P., 2000. Modelling changes
in soil organic matter after planting fast-growing Pinus radiata on Mediterranean
agricultural soils. Eur. J. Soil Sci. 51, 627–641.

20. Romero-Díaz, A., Ruiz-Sinoga, J.D., Robledano-Aymerich, F., Brevik, E.C., Cerdà,
A., 2017. Ecosystem responses to land abandonment in Western Mediterranean
Mountains. Catena. https://doi.org/10.1016/j.catena.2016.08.013

21. Ruecker, G., Schad, P., Alcubilla, M.M., Ferrer, C., 1998. Natural regeneration of
degraded soils and site changes on abandoned agricultural terraces in Mediterranean
Spain. L. Degrad. Dev. 9, 179–188.

22. Valverde-Asenjo, I., Diéguez-Antón, A., Martín-Sanz, J.P., Molina, J.A., Quintana,
J.R., 2020. Soil and vegetation dynamics in a chronosequence of abandoned vineyards.
Agric. Ecosyst. Environ. 301. https://doi.org/10.1016/j.agee.2020.107049

102
CHAPTER III: Soil organic carbon accumulation rates on Mediterranean abandoned agricultural lands

23. Zethof, J.H.T., Cammeraat, E.L.H., Nadal-Romero, E., 2019. The enhancing effect of
afforestation over secondary succession on soil quality under semiarid climate
conditions. Sci. Total Environ. 652, 1090–1101.
https://doi.org/10.1016/J.SCITOTENV.2018.10.235

24. Zornoza, R., Mataix-Solera, J., Guerrero, C., Arcenegui, V., Mataix-Beneyto, J., 2009.
Comparison of soil physical, chemical, and biochemical properties among native forest,
maintained and abandoned almond orchards in mountainous areas of Eastern Spain.
Arid L. Res. Manag. 23, 267–282. https://doi.org/10.1080/15324980903231868

103
CHAPTER IV: Factors driving soil carbon
sequestration following agricultural land
abandonment in Europe
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

4.1 Overview
Agricultural land abandonment (ALA) is a prominent land use change throughout Europe, with
several notable implications for soil health and ecosystem restoration. In particular, the
cessation of intensive agricultural practices often induces an increase in soil organic carbon
(SOC) and can potentially support land-based climate change mitigation efforts. However,
large uncertainties on the variability of post-abandonment soil carbon sequestration (SCS) rates
and the absolute storage potentials across Europe hinders the development of dedicated policies
leveraging the ecological benefits of both planned and unplanned ALA. In this chapter, I
collected and synthesized SOC stock changes following ALA derived from field sites in
European countries using published chronosequence/paired plot data (804 observations, 546
soil profiles). In doing so, I determined how rates of soil carbon accumulation during ecological
succession differ in space and time. I found a slow, but significant, rate of SOC stock increase
across Europe of 1.28% yr–1, and an absolute rate of 0.32 Mg ha–1 yr–1. The average relative
and absolute changes are +32.1% and +10.5 Mg ha–1, respectively, with an average time since
abandonment of 34 years. SOC responses were negatively correlated with initial SOC stock,
indicating a soil carbon saturation effect. Low initial stock (< 25 Mg ha–1 at 0–30 cm depth)
exhibited a significantly higher SCS rate than all the other initial stock classes, accumulating
SOC at 1.95% yr–1. Abandoned agricultural lands in biogeographical regions featuring optimal
climatic windows showed greater SOC sequestration rates, with mean soil carbon change
ranked from highest to lowest as Pannonian > Mediterranean > Atlantic > Continental > Boreal
> Alpine. However, climatic conditions and human management factors can have both positive
and negative effects on SOC, resulting in several strongly divergent responses to ALA. Past
croplands had a notably greater rate of SOC increase over time relatively (1.52% yr–1) and
absolutely (0.38 Mg ha–1 yr–1) than sites that were previously used as pastures, likely a result
of lower initial SOC stocks in croplands compared to pastures. Sites that underwent natural
ecological succession exhibited a greater rate of change in SOC stock relatively (1.59% yr–1)
and absolutely (0.35 Mg ha–1 yr–1) compared to sites that were actively restored or converted
to new vegetation land covers, for example through tree planting practices. These results
provide some clarity on previous regional debates surrounding the positive, negative, and
neutral SCS potential of post-agricultural soils, which have likely been confounded by the
factors investigated here. Abandoned croplands with low initial SOC stock and managed
through natural succession can be expected to show the greatest SOC accrual in Europe, while
fertile pastures that are actively converted (e.g., afforested) would result in the lowest increases
in SOC, or even losses. This variability in post-abandonment/conversion SOC dynamics must
be considered in sustainable land use planning that strives to incorporate the positive ecological
and climate change mitigation implications of ALA, taking into account site-specific
conditions and past and present land management regimes to avoid negative impacts for soil
health and lost opportunities for climate change mitigation.

106
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

4.2 Introduction
Agricultural land abandonment (ALA) is a prominent global land use change. But despite its
ubiquity across agricultural regions (Campbell et al., 2008; Li and Li, 2017), it is difficult to
accurately measure and monitor at large geographic scales, which can lead to large
uncertainties (Yin et al., 2020). This is due to both its multiple definitions as a land use change
and its rather ephemeral nature as a land use classification, often undergoing recultivation after
short periods (e.g., overlapping with shifting agriculture or unreported and informal fallowing
practices) (Benjamin et al., 2007; Heinimann et al., 2017). The lack of incentives and interest
to report genuine ALA also produces inaccurate land use inventories and low-resolution
mapping in under-resourced regions. In the European Union however, efforts to monitor,
measure, and map ALA have achieved a comparatively higher level of success due to the
incorporation of multiple sources of predictive variables and model parameters (i.e., the LUISA
modelling platform, see Lavalle et al., (2020)). These efforts suggest that more than 5.6 Mha
is predicted to be abandoned in the EU and the UK by 2030, or 3.6% of total agricultural land
(Perpiña Castillo et al., 2021). Indeed, most of the +1.4% total forest area increase across the
continent from 1992 to 2015 has occurred on former agricultural lands (Palmero-Iniesta et al.,
2021).

In non-degraded agricultural landscapes, the cessation of intensive agricultural practices


typically results in the spontaneous recovery of ecosystem properties towards pre-agricultural
levels (Cramer et al., 2008). Ecosystem health indicators for vegetation, soil, and wildlife can
all improve from the ensuing ecological succession following ALA. These trajectories depend
significantly on site-specific conditions (e.g., the level of degradation, the suitability for
restoration, and the level of biodiversity one would find there naturally compared to the level
of biodiversity maintained by the active agroecosystem before ALA) (Beilin et al., 2014;
Plieninger et al., 2014; Queiroz et al., 2014). Although conventional agricultural practices are
known to continuously deplete soil carbon (Carlson et al., 2017; Lal, 2013), one of the most
important benefits of the natural recovery of post-agricultural soils is the regeneration of soil
carbon stocks (Deng et al., 2014; Laganière et al., 2010; Wertebach et al., 2017).

As the largest terrestrial carbon pool that can be effectively influenced by human efforts, soil
organic carbon (SOC) represents a critical carbon sink for climate change mitigation efforts
(Lal, 2004). The natural ability of post-agricultural soils to reabsorb carbon until, presumably,
reaching pre-agricultural levels has therefore received increasing attention, especially for its
implications for sustainable land management (CHAPTER II; Lasanta et al., 2015; Schröder et

107
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

al., 2018). Unfortunately, there remains several uncertainties surrounding ALA’s potential as
a climate change mitigation tool via soil carbon sequestration (SCS) at a regional, continental,
and global scale, despite notable instances in history of large-scale sequestration (e.g.,
following the collapse of the former Soviet Union (Kuemmerle et al., 2011; Schierhorn et al.,
2013; Wertebach et al., 2017), or even following the mass die-off of pre-colonial South
America (Koch et al., 2019)). Not all landscapes accumulate carbon at the same intensity (i.e.,
amount of stock increase) and speeds (i.e., rate of stock increase) following ALA (Breuer et
al., 2006; Hoogmoed et al., 2012; Nadal-Romero et al., 2016), and under specific conditions
some can even lose soil carbon (Martinez-Duro et al., 2010; Segura et al., 2020).

In general, new policies on sustainable land management are expected to include stipulations
for protecting and replenishing SOC stocks whenever possible (Amelung et al., 2020; Bossio
et al., 2020; Bradford et al., 2019). The ability to calculate the rates and amounts of post-
agricultural SCS across large geographies is pivotal for the planning, implementation, and
assessment of land management policies that incorporate ecosystem restoration aspects (Xie et
al., 2020). This is made even more necessary by the fact that ALA is a continuous (i.e., both
historically and currently relevant) and often unplanned land use change (LUC) that is already
influencing soil carbon stocks, for better or worse, and needs to be more accurately quantified.
Therefore, the combination of robust datasets that produce region- and management-specific
rates of SCS with accurate and detailed maps of ALA in that region (i.e., the spatial body on
which to apply the rates) creates the possibility to support effective and climate-smart land
management policies (Cook-Patton et al., 2020; Vermeulen et al., 2019).

Europe, where 11% of total greenhouse gas emissions stem from agriculture (EU NIR, 2021),
represents an ideal combination of data resources and land use histories for an integrated, large-
scale study on the SCS potential and implications of ALA. This is due to the widespread
historical and ongoing ALA (Estel et al., 2015; Lasanta et al., 2017; Levers et al., 2018;
Ustaoglu and Collier, 2018), the recent political and scientific push for effective, efficient, and
accessible SCS strategies (Gardi et al., 2016; Montanarella and Panagos, 2021; Navarro and
Pereira, 2012; Rodrigues et al., 2021; Schröder et al., 2018); the availability of published
studies on soil properties and ALA (i.e., chronosequence and paired-plot data); and the detailed,
robust, and up-to-date land use/cover inventory keeping that has produced high-quality spatial
projections of ALA (Lavalle et al., 2020; Perpiña Castillo et al., 2021). The continental
coverage of published chronosequence and paired-plot data, in particular, allows for the
quantification of total soil carbon stock changes following ALA, the determination of cluster-

108
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

specific SCS rates, and the elucidation of the modulating factors on post-agricultural SCS.
Despite the increasing focus on this topic in recent years, there remains much uncertainty on
the direction of soil carbon response to ALA (i.e., increase, decrease, or no change), the
intensity (i.e., how much change), and the duration (i.e., how long will the change last) in the
various biogeographical regions of the EU and its neighbouring countries, likely due to the
complex interactions of the modulating factors and their confounding effects. Afterall, the EU’s
Thematic Strategy for Soil Protection has listed SOC protection and enhancement as a
necessary goal for member states, and SOC loss as one of the eight soil threats on the continent
(EC, 2012).

In light of these challenges and opportunities for both research advancement and policy
support, here I synthesized published chronosequence and paired-plot data from field sites
within Europe and explored the variability in SOC responses to ALA and direct conversion
from agriculture to re-naturalized landscapes. I conducted a literature search to combine all
previously synthesized data at different geographical scales with never-before synthesized
published studies and I categorized each chronosequence/paired-plot collected based on several
key factors that may influence SOC stock dynamics (i.e., climate, biogeographical region, past
land use, past crop type, and present land management). I expect a noticeable increase in SOC
across Europe following abandonment/conversion, but with high variability in sequestration
rates. These results are intended to provide important context on the soil carbon implications
of land use change in Europe, particularly from an ecosystem restoration perspective.

4.3 Methods
4.3.1 Literature search and data collection
Published chronosequence and paired-plot studies undertaken in Europe investigating the
impacts of ALA (or direct land use conversions from agriculture) on grassland, shrubland, and
forest succession were compiled for analysis following a literature search. While repeated
measurements are the most ideal approach for determining the effects of land use change over
time, chronosequences and paired-plots are proven alternatives commonly employed (Breuer
et al., 2006; Walker et al., 2010).

The literature search and data collection process comprised of two stages. In stage one, an
initial dataset of relevant studies was established by identifying any individual studies that
included European sampling sites from the databases and lists of references of previously
published synthesis studies on thematically related topics (i.e., land use changes and soil
properties) at any geographic scale (i.e., regional syntheses within Europe, syntheses of Europe,

109
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

and global syntheses including Europe) (Table 9). This allowed me to collate all previously
synthesized, time-stamped, post-agricultural SOC data in Europe into one combined dataset.

Table 9. Results of the first stage of data collection process. Previously published synthesis
studies thematically related to the present study (i.e., land use changes and soil properties) at
any geographic scale (i.e., regional syntheses within Europe, syntheses of Europe, and global
syntheses including Europe) were surveyed for relevant individual studies. Relevant synthesis
studies that were found to contain zero relevant individual studies are not included (e.g., Guo
and Gifford, (2002)).

Synthesis study Spatial Number of Final number Final number


extent relevant of relevant of data-pairs
individual individual extracted
studies studies from
identified retained after remaining
within exclusion individual
criteria

Li et al., (2018) Global 85 4 66

Deng et al., (2016) Global 49 3 21

Bárcena et al., (2014) Europe 2 1 21


(Northern)

Li et al., (2012) Global 49 1 9

Post and Kwon, (2000) Global 7 0 0

Laganière et al., (2010) Global 6 0 0

Paul et al., (2002) Global 13 0 0

Conant et al., (2001) Global 3 0 0

Shi et al., (2013) Global 5 0 0

Kämpf et al., (2016) Temperate 12 0 0

Poeplau et al., (2011) Temperate 11 0 0

TOTAL 242 9 117

The second stage of the literature search targeted all new and/or previously un-synthesized
individual studies with relevant data. The following key terms were searched in November
2020 using ISI Web of Science with results limited to English language studies published in
any year: (plough* OR till* OR crop* OR farm* OR agri* OR cultivat* OR *field OR pasture
OR meadow OR grazing OR range*) AND (*forest* OR grassland OR shrubland OR natural

110
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

OR secondary OR recover* OR plantation* OR conver* OR abandon* OR old* OR regenerat*


OR *aside OR restor* OR succession* OR fallow OR revegetat*) AND (chronosequence OR
pair*) AND (soil OR carbon). This initial search resulted in 4718 hits, which were then sorted
by geographic region, producing a subset of studies that had related terms either in their title,
abstract, or keywords (i.e., “Europe” related terms, European country names, and geographic
feature names that indicated potential sampling sites in Europe (e.g., Alps, Nordic,
Mediterranean)). Another ISI Web of Science search was conducted in early 2022, following
the same procedure, to update the dataset with relevant studies published since the original
search date in 2020. In addition, relevant studies discovered through snowballing reference lists
and “recently cited by” lists were included to further supplement the dataset.

The final list of relevant papers from these two stages of the literature search were then
subjected to inclusion/exclusion criteria before data extraction. For an individual study to be
included, the time since abandonment/conversion (years) and the SOC or soil organic matter
(SOM) concentration or stock (various units) of the mineral soil at any depth for
chronosequence stage or paired-plot must have been provided. Each chronosequence and
paired-plot must have featured one agricultural control field (i.e., actively cultivated,
representing 0 years since abandonment/conversion) to compare the treatment field(s) to (i.e.,
abandoned or converted from agriculture). The following secondary criteria for both the control
and treatment fields were either extracted from the studies themselves, provided by authors
upon request, or determined through either online sources or inferred empirically during data
processing (see below): sample bulk density (BD), sampling site coordinates (latitude,
longitude), mean annual precipitation (mm), mean annual temperature (° C), past land use
(cropland, pasture), past crop type (woody, annual), post-ALA/conversion land management
system (natural, assisted, occasionally grazed), vegetation type restored (forest, shrubland,
grassland), and biogeographical region as per the European Environment Agency classification
system (EEA, 2016), and the soil sample depth (upper and lower), sample size (number of
samples), and error estimates for SOC/SOM/bulk density values reported (standard deviation
or standard error). Studies were excluded if they were in locations outside of the
biogeographical region coverage of Europe, or if they failed to provide a means to determine
any of the previously outlined criteria.

Soil carbon concentration and stock data in either SOC or SOM were extracted from tables,
text, supplementary files, graphs/figures by digitizing (GetData Graph Digitizer, v.2.26,
Russia), or by request to the corresponding authors. To ensure both accuracy and comparability

111
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

of the data, all values were collected directly from the original published source or author, and
never from secondary datasets contained in the synthesis studies listed in Table 9.

Figure 18. Distribution of chronosequence/paired-plot data-pairs in the dataset according to


the year of publication of the original studies.
The final dataset from both stages of the literature and data collection process featured studies
from most of the EU27 member states, in addition to the United Kingdom (UK), Switzerland
(CH) and Norway (NO). A total of 102 studies published from 1994 to 2022 were identified
under the inclusion/exclusion criteria (Figure 18, Supplementary materials Table 12),
representing 804 time-stamped data-pairs of control and abandoned/converted soils sampled
throughout the EU27+UK+CH+NO (Figure 19) at any depth. The first stage of data collection
resulted in 117 data-pairs (Table 9), while the second stage resulted in 687. The dataset ranges
in time since abandonment/conversion from 1 to 193 years.

112
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

Figure 19. Map of the sampling sites of the 804 data-pairs from the 102 published studies
located within the extent of the Biogeographical Regions of Europe (EEA, 2016).
4.3.2 Data processing and analysis
Within the full dataset, there were 438 data-pairs with SOC stock already reported and collected
(54% of the dataset). To calculate the SOC stock of the remaining 366 data-pairs, several steps
were needed. When only SOM data was provided (n = 89), SOC concentration values were
calculated from the revised Van Bemmelen conversion factor (0.5 instead of 0.58) following
Cook-Patton et al., (2020), based on Pribyl, (2010) using Eq. (1):

𝑆𝑂𝐶 = 𝑆𝑂𝑀 × 0.5 (1)

All SOC concentration values for the 366 data-pairs were then standardized to the same
concentration unit (%). To calculate the SOC stock from these values, BD is necessary and was
already reported for 134 of the remaining data-pairs. When BD was not reported (n = 232), it
was estimated based on the available SOC data according to the pedotransfer function of
Manrique and Jones, (1991) shown in Eq. (2):

113
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

𝐵𝐷𝑒 = 1.660 − 0.318(𝑆𝑂𝐶𝐶 )1/2 (2)

where BDe is the estimated bulk density (g cm–3) and SOCc is the reported SOC concentration
(%). The reported and estimated BD values from Eq. (2) where then used to calculate SOC
stocks in Mg ha–1 at the respective sampling depth for each of the 366 data-pairs according to
Eq. (3):

𝑆𝑂𝐶𝑠𝑡 = 𝑆𝑂𝐶𝑐 × 𝐵𝐷𝑚/𝑒 × 𝐷 (3)

where SOCst is the SOC stock (Mg ha–1), SOCc is the SOC concentration (%), BDm/e is the
measured or estimated bulk density (g cm–3), and D is the soil depth sampled (cm) based on
the upper and lower depth of the sample. Because not all of the studies reported BD for each
of the soil depths sampled, it was not possible to correct the entire dataset to a common soil
mass, although this limitation is generally accepted in large-scale syntheses as it is not expected
to result in a significant bias in SOC stock estimates following land use change (Deng et al.,
2016; Guo and Gifford, 2002; Laganière et al., 2010).

The full dataset (n = 804) of reported and calculated SOC stocks at all depths was then
standardized to a depth of 30 cm to increase the comparability of the study sites. Each soil
profile was combined to produce a total stock per profile at a known maximum depth from the
surface, reducing the dataset to 546 data-pairs. The SOC stock in the top-30 cm of each profile
was then estimated following the methodology of Deng et al., (2016), using the SOC depth
distribution function developed by Jobbágy and Jackson, (2000), according to Eq. (4) and (5):

𝑌 = 1 − 𝛽𝑑 (4)

1 − 𝛽 30 (5)
𝑋30 = × 𝑋𝑑0
1 − 𝛽 𝑑0

where Y is the cumulative proportion of the SOC stock from the soil surface to depth d (cm); β
is the relative rate of decrease in the SOC stock with soil depth; X30 denotes the SOC stock in
the upper 30 cm; d0 denotes the original soil depth from the single or combined soil profile
(cm); and Xd0 is the original SOC stock from the single or combine soil profile (Mg ha–1). The
lack of significant differences detected between SOC depth distribution functions analysed in
global biomes in Jobbágy and Jackson, (2000), and the large biogeographic scale of the present
study, allows for the adoption of the global average SOC depth distribution β value (0.9786),

114
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

which is commonly employed in studies with similar approaches, data, and large spatial scales
(Deng et al., 2016; Li et al., 2012). Although this estimation introduces uncertainties in the data
accuracy (e.g., the fact that control and treatment sites undoubtably have different SOC depth
distributions in reality, as well as the treatment sites at different successional stages (see, for
example, CHAPTER II, Figure 13)), it is not expected to skew overall trends of SOC dynamics
during revegetation when analysing this kind of data at large geographic scales for the
determination of generalized land use change effects (Li et al., 2012; Yang et al., 2011).

To standardize the effect of time since abandonment/conversion on SOC stocks between the
various chronosequences and paired-plots of all the studies at the 30 cm depth, all profiles were
plotted as the absolute and relative difference from their paired agricultural control SOC values,
according to Eq. (5) and (6):

∆𝑆𝑂𝐶𝑠𝑡_𝑎𝑏𝑠 = 𝑆𝑂𝐶𝑠𝑡_𝑝𝑜𝑠𝑡 − 𝑆𝑂𝐶𝑠𝑡_𝑝𝑟𝑒 (5)

𝑆𝑂𝐶𝑠𝑡_𝑝𝑜𝑠𝑡 − 𝑆𝑂𝐶𝑠𝑡_𝑝𝑟𝑒 (6)


∆𝑆𝑂𝐶𝑠𝑡_𝑟𝑒𝑙 = × 100
𝑆𝑂𝐶𝑠𝑡_𝑝𝑟𝑒

where SOCst_rel is the relative change in SOC stock (%), SOCst_abs is the absolute change in
SOC stock (Mg ha–1), SOCst_post is the SOC stock after abandonment/conversion (Mg ha–1) (i.e.,
the treatment), and SOCst_pre is the SOC stock before abandonment/conversion (Mg ha–1) (i.e.,
the control). The relative change in SOC stock data were fit to linear regressions with 95%
confidence intervals to determine the general directional responses of SOC to time since
abandonment considering various climatic factors, biogeographical regions, past land uses, past
crop types, and management factors (assuming significance at p < .05 using Grapher (Golden
Software, v.15, USA)). By dividing the SOCst_abs or SOCst_rel by the age (time since
abandonment), the absolute and relative SOC sequestration rates can be determined for each
soil profile at its respective point in time, while the slope of the linear regressions provide more
generalized SOC sequestration/loss rates for the entire category of data modelled. The overall
changes in SOC stock (sequestration or loss) for each of the main categories within each overall
factor, irrespective of time, were summarized for simplified comparison in forest plots with
95% confidence intervals, based on Eq. (7) and (8):

(7)
𝑉𝑠
𝑆𝐸𝑡𝑜𝑡𝑎𝑙 = √
𝑁

115
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

95% 𝐶𝐼 = 1.96 × 𝑆𝐸𝑡𝑜𝑡𝑎𝑙 (8)

Where SEtotal is the standard error of the change in SOC stocks per category, Vs is the variance,
and N is the number of samples. Due to a Spearman rank correlation test revealing a weak
significant correlation between sample size and effect size (rho = –0.08, p = 0.04) in the full
profile dataset (n = 804) using R statistical software (R Core Team, 2020), a funnel plot was
used to graphically inspect the possibility of skewed data (Figure 20).

Figure 20. Funnel plot of the full dataset (n = 804)


The presence of publication bias is considered high if the funnel shape is irregular; for example,
with the narrow end of the funnel closer to the y-axis than the wide end of the funnel. Figure
20 indicates low likelihood of publication bias, with a regular funnel distribution as seen in
other large SOC datasets built from 50+ individual studies (see, for example, Figure 1 in Kämpf
et al., (2016)). The impact of each predictor variable in the dataset on SCS, and all their
potential pairwise interactions, was explored through the R package ‘gmulti’ (Calcagno and
Mazancourt, 2010), whereby the best fitting generalized linear model was determined
(candidate screening parameters set to “exhaustive”, and ranking criteria set to Akaike
information criterion (AICc)). The best fitting model’s performance was not improved when
run as a generalized additive model (Gaussian), and was therefore assumed to be linear. As
analysis of variance (ANOVA) requirements for data distribution normality were not met

116
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

(Shapiro–Wilk test), non-parametric Kruskal–Wallis rank sum tests were also performed in R
statistical software to further test for significant differences of the factor effects on soil carbon
changes (significance at p < 0.05) (R Core Team, 2020).

4.4 Results & Discussion


4.4.1 Overall factors driving soil carbon sequestration following ALA
The average absolute SOC change amongst the 546 data-pairs is 10.5±2.48 Mg ha–1, with an
average time since abandonment of 32 years. Each of the groups of factors contained significant
differences between at least one pair of categories (Supplementary materials Table 10). The
overall SOC stock trends for each of the dominant non-human related factors (i.e., site-specific)
and human-related factors (i.e., management) are illustrated in Figure 21 and Figure 22,
respectively. The data-pairs were classified based on their time since abandonment/conversion
as age classes being either young (≤ 10 years), early-stage (> 10 to ≤ 20 years), middle-stage
(> 20 to ≤ 40 years), or late-stage (> 40 years) succession. All stages have positive mean SCS
values, and the overall trend was as expected, with young sites having the lowest increase
(3.07±2.66 Mg ha–1, n = 105), followed by early-stage (4.71±4.56 Mg ha–1, n = 108), middle-
stage (9.66±4.53 Mg ha–1, n = 153) and late-stage (19.07±5.41 Mg ha–1, n = 180) sites. In many
cases, SOC can be lost in the first ~5–10 years following abandonment/conversion from
agriculture as vegetation regrowth begins and SOC remains in flux, followed by a return to pre-
abandonment/conversion levels and then a net increase in the following decades (Deng et al.,
2014). In cases of afforestation of abandoned agricultural lands, Paul et al., (2002) identified a
3–35 year period of initial decrease, after which net SCS can be expected from ~30 years. An
initial period of flux can be expected as both SOM stabilizing (e.g., manure and organic
amendment application) and SOM disrupting (e.g., ploughing, crop residue removal) practices
cease. This implies the importance of time since abandonment generally, however the
unclassified and independent effect of time is less clear (e.g., Supplementary materials Figure
28).

117
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

Figure 21. Effects of dominant non-human-related factors on soil carbon change (Mg ha–1)
following abandonment/conversion. Age classes are young (≤ 10 years), early-stage (> 10 to
≤ 20 years), middle-stage (> 20 to ≤ 40 years), or late-stage (> 40 years) successional sites.
Symbols and error bars indicate mean ±95% CI. Dashed line indicates x = 0. Numbers in
parentheses indicate number of observations.
Nearly all categories showed a significantly positive overall effect on SOC change. The only
categories with insignificant changes in this dataset, based on 95% confidence intervals, were
sites with > 900 mm MAP (2.73±4.91 Mg ha–1, n = 184), < 8 ° C MAT (3.63±4.21 Mg ha–1, n
= 171), and the associated Biogeographical regions closer to these climatic thresholds: Atlantic
(8.86±8.95 Mg ha–1, n = 64), Boreal (1.09±9.43 Mg ha–1, n = 10), and Alpine (-7.25±6.66 Mg
ha–1, n = 80). The effects of the initial SOC stock at the time of abandonment/conversion are
also clearly evident in Figure 21, exhibiting the expected inverse relationship with SOC change.
The lowest initial stock category (< 25 Mg ha–1, n = 91) responded the greatest to
abandonment/conversion, increasing by an average of 22.34±4.73 Mg ha–1, followed by
increasingly larger initial stocks categories of > 25-50, > 50-75, and > 75 Mg ha–1, increasing
by 9.98±2.38 (n = 124), 9±5.22 (n = 106), 6.74±4.86 Mg ha–1 (n = 225), respectively.

118
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

Figure 22. Effects of dominant human-related factors on soil carbon change (Mg ha–1)
following abandonment/conversion. Symbols and error bars indicate mean ±95% CI. Dashed
line indicates x = 0. Numbers in parentheses indicate number of observations.
The categories within the human-related factors also exhibited several expected trends
identified in previous studies covering different geographic extents (Deng et al., 2016; Kämpf
et al., 2016). Although all past croplands exhibited a positive increase (14.75±2.62 Mg ha–1, n
= 395), there is a significant difference between past woody croplands (e.g., vineyards, olives
groves, orchards) at 25.88±4.5 Mg ha–1 (n = 126) and past annual croplands (e.g., cereals) at
9.31±3.11 Mg ha–1 (n = 261). On the other hand, lands that were previously used as pastures
did not demonstrate a significant response to abandonment/conversion (-0.56±5.38 Mg ha–1, n
= 151). Accordingly, all land use and land cover change categories involving
abandonment/conversion of pastures to more naturalized vegetation communities (i.e.,
grassland, shrubland, or forest) did not exhibit statistically significant SOC responses, while
all categories involving croplands produced significant positive responses in SOC, ranging
from 12.64–23.43 Mg ha–1. The categories of land management regimes post-ALA/conversion,
interestingly, did not differ as significantly as the categories within the other driving factors
(MGMT in Supplementary materials Table 10). However, natural succession (n = 326), also

119
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

known as passive management, did result in a higher overall soil carbon change than assisted
restoration (n = 194), also known as active management (12.22±3.05 Mg ha–1 compared to
8.69±4.65 Mg ha–1).

4.4.2 Overall SOC dynamics following ALA


Despite notable variability in responses, abandonment/conversion from agricultural practices
across Europe results in a slow, but significant, relative rate of SOC stock increase of 1.28%
yr–1 (n = 546, R2 = 0.19, p < 0.0001) (Figure 23.a) and an absolute rate of 0.32 Mg ha–1 yr–1 (n
= 546, R2 = 0.09, p < 0.0001) (Figure 23.c). This absolute rate is quite comparable to other
large geographic scale studies, like Post and Kwon (2000) who found global croplands to forest
conversions sequestered 0.34 Mg ha–1 yr–1 (n = 47), and Deng et al., (2016) who calculated a
sequestration rate of 0.30 Mg ha–1 yr–1 for global croplands to grassland conversions (n = 57).
Deng et al., (2014) also reported another highly similar rate of 0.33 Mg ha–1 yr–1 based on 844
observations at 181 sites from China’s “Grain-for-Green” Program. However, at the biome and
regional scales, several other synthesis studies have reported larger rates (see Figure 10 in
CHAPTER II). For arable land to managed or unmanaged grasslands conversions in the
temperate zone, Kämpf et al., (2016) found a SCS rate of 0.72 Mg ha–1 yr–1 (n = 54), which
may be overestimated due to the comparatively shorter average time since
abandonment/conversion (14 years). SCS rates even as high as 1.30 Mg C ha−1 yr−1 have been
estimated across the tropical zone and in China (Deng et al., 2014; Silver et al., 2001). In
another example, the weighted average rate of two studies exploring the SCS potential of active
restoration on at the global scale is 0.87 Mg C ha−1 yr−1 (Li et al., 2012; Shi et al., 2013). It is
important to remember that directly comparing reported SCS rates between synthesis studies
is challenging due to differences in sample site distribution and study parameters (i.e., time
range since abandonment/conversion, soil depth considered, management practices included
or excluded, etc.).

On a logarithmic scale, the positive correlation between time and SOC stock change is more
noticeable, reaching a clearer direct relationship at the X,Y extremes (i.e., arrowhead
converging to a 1:1 correlation at longer time scales) (Figure 23.b,d). These results provide
new insight into some of the previous regional debates on the positive, negative, neutral SCS
potential of post-agricultural soils (Bárcena et al., 2014; CHAPTER III), which have likely
been confounded by other key factors examined here. The model-average importance of each
predictor factor explored is summarized in Supplementary materials Figure 28 with the results
of the multi-model analysis and best model equation provided in Supplementary materials

120
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

Table 11. Time since abandonment alone was not among the most important predictor factors,
indicating the highly complex interactions at-play following abandonment/conversion, at least
over the distribution of time periods available in this dataset (Q1 of 14 years, median of 27
years, mean of 32 years, and Q3 of 41 years). The overwhelming variability of SOC responses
in the first several decades demonstrates the importance of considering all potential factors in
addition to the time-scale reported. Long-term land management scenarios must therefore be
detailed enough with these factors to adequately capture the uncertainty in SOC responses,
especially in the first few decades. The distribution of time since abandonment in this dataset
may not be sufficient to fully constrain the isolated effect of time over stronger cofounding
interactive effects of human management and environmental factors.

121
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

Figure 23. Relative and absolute change in SOC stock over time since
abandonment/conversion (yr) on a linear scale with regression results shown in the insert (a,
c) and on a logarithmic scale (b, e). Relative change in SOC stock (%) against initial SOC
stock (Mg ha–1) of the full non-standardized dataset (n = 804) on a linear scale (e) and on a
logarithmic scale (f). Shaded area represents 95% confidence interval.

122
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

The influence of the initial SOC stock at the time of abandonment/conversion is also evident
in Figure 23. The ability of post-agricultural soils with high initial SOC stocks to accrue new
carbon were limited, whereas the soils with the highest relative increases in SOC stock were
exclusively ones that had very low initial SOC stock (Figure 23.e). However, many soils with
low initial SOC stock also had very low or even negative SOC responses to
abandonment/conversion. Overall, the relationship between initial SOC stock and relative SOC
increase is negative (Figure 21, Figure 23.f). This relationship is to be expected based on
classical soil carbon saturation theory, with the soils with greater initial stock presumably
closer to their saturation limit and therefore with less capacity to accrue new SOC (Stewart et
al., 2007). Amongst the different soil depth classifications examined in the full dataset (n =
804), the relative change in SOC stock followed expected patterns (Figure 24.b). The highest
rates of increase were found in the top-soil at < 5 cm (2.52% yr–1, n = 79, R2 = 0.51, p < 0.0001),
followed by the 5-15 cm depth (1.00% yr–1, n = 366, R2 = 0.16, p < 0.0001). The > 15 cm depth
exhibited no statistically significant change over time (n = 359, p = 0.436), with the potential
of SOC losses after several decades. The effect of initial SOC stock is also present when
combined with depth (Figure 24.a) with higher initial stock soils (> 50 Mg ha–1) exhibiting
slower rates of SOC change than lower initial stock (< 50 Mg ha–1) soils for each maximum
depth grouping of complete soil profiles measured (0–10 cm and 0–30 cm).

123
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

Figure 24. (a) Relative change in SOC stock (%) over time since abandonment/conversion (yr)
for different initial SOC stocks found in complete soil profiles with different maximum soil
depths (0–10 or 0–30), representing a subset of the total dataset. L indicates low initial stock
(< 50 Mg ha–1) and H indicates high initial stock (> 50 Mg ha–1). Shaded areas represent
respective 95% confidence intervals of linear regressions. (b) Relative change in SOC stock
(%) over time since abandonment/conversion (yr) for different soil depths (average) within the
total dataset (n = 804). (c, d) Relative and absolute change in SOC stock of different initial
SOC stock classes (Mg ha–1) in the standardized dataset (n = 546), plotted over time since
abandonment/conversion (yr) on a linear scale. Numbers in parenthesis indicate sample sizes.
Soil nutrients accumulation is known to be highest closer to the surface during secondary
succession, where plant litter inputs are present and there is comparatively more biochemical
processes and exchanges occurring (Cramer et al., 2008; Hu et al., 2018; La Mantia et al., 2013;
Nadal-Romero et al., 2016). Although post-agricultural soil profiles may demonstrate a distinct
legacy of tillage in having a lasting homogeneity (e.g., Sulman et al., (2020)), it is also possible
that the ability of new SOC to saturate deeper into the previously homogenized and SOC-
depleted top/mid-soils is outpaced by the accrual of new SOC at the surface resulting in greater

124
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

rates of SOC increases over time (CHAPTER III, Figure 13). In absolute terms across the
standardized dataset (n = 546), the effects of initial stock over time since abandonment are less
clear than in relative terms (Figure 24.c,d). Low initial stock (< 25 Mg ha–1) exhibited a
significantly higher SCS rate than all the other initial stock classes (1.95% yr–1, n = 91, R2 =
0.41, p < 0.0001), while the highest initial stock (> 75 Mg ha–1) exhibited a significantly lower
sequestration rate than all the other classes (0.22% yr–1, n = 225, R2 = 0.03, p < 0.007).

4.4.2.1 Climatic and biogeographical factors on SOC dynamics


The climatic regime present at the sampling sites had a noticeable influence on the rates of
SOC change following abandonment/conversion from agricultural practices. Similar to the
results of a synthesis of natural succession post-agricultural chronosequences and paired-plots
in peninsular Spain (CHAPTER III), the sampling sites distributed across Europe were subject
to complex temperature (Figure 25.top) and precipitation (Figure 25.bottom) windows for post-
agricultural SOC accumulation.

125
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

Figure 25. Absolute (Mg ha–1) and relative (%) change in SOC stock over time since
abandonment/conversion (yr) for mean annual temperature (MAT, C, top panels) and mean
annual precipitation (MAP, mm, bottom panels) and their linear regressions. Shaded areas
represent respective 95% confidence intervals. Numbers in parenthesis indicate sample sizes.
SOC stock change depends on high precipitation and temperature for organic matter input
through the increased net primary productivity. However, in many regions of Europe that is
represented in this dataset, confounding climatic effects are present (e.g., cold and wet Alpine
pastures, hot and dry Mediterranean vineyards). This climatic complexity in SCS following
ALA can also be seen through the lens of biogeographical region classifications, which take
into account specific vegetation, biodiversity, and climate interactions (Figure 26), and are
linked closely with soil processes (Ibáñez et al., 2013). Interestingly, while cold and wet
climates are normally associated with higher SOC accumulation for most land use changes, it

126
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

may not hold true for ALA (Gardi et al, 2016; Guo and Gifford, 2002). For ALA, the apparent
limiting effect of precipitation at levels above 900–1100 mm per year that has been reported
across the Mediterranean is likely the result of precipitation induced N leaching, decreases in
aggregate protected SOC, and increases in less protected particulate SOM fractions (Alberti et
al., 2011; Gabarrón-Galeote et al., 2015b; Guidi et al., 2014; Navas et al., 2012). At the global
scale, precipitation and SOC accumulation during ecological succession generally correlate
negatively (Jackson et al., 2002), although dry conditions found in semi-arid climates (e.g., <
450 mm yr–1) can also limit net primary productivity (NPP) and therefore limit organic matter
inputs that promote SOC accumulation (Bonet, 2004; Gabarrón-Galeote et al., 2015b;
Robledano-Aymerich et al., 2014). Conversely, while increased soil carbon cycling can be
linked with increasing temperature, high SOM decomposition rates limiting SCS in sermi-arid
climates are unlikely because microbial activity would also be drought limited (Moreno et al.,
2019). For grasslands on previously managed pastures, SOC also correlates positively with
temperature and negatively with precipitation (Kämpf et al., 2016; La Mantia et al., 2013; Pellis
et al., 2019). As a partial representation of climatic conditions, the European biogeographical
regions also display a wide variety of rates of change for SOC stock following
abandonment/conversion (Figure 26). These biogeographical regions are based on biota, unlike
biomes, and emphasize endemic and/or spatially distinct and limited taxa and communities
(Morrone, 2018).

Figure 26. Absolute (left, Mg ha–1) and relative (right, %) change in SOC stock (%) over time
since abandonment/conversion (yr) for the Biogeographical Regions of Europe and their linear
regressions. Shaded areas represent respective 95% confidence intervals. Numbers in
parenthesis indicate sample sizes.

127
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

Some of the highest relative rates of SOC stock increase were from soils within the
Mediterranean biogeographical region (1.26% yr–1, n = 187, R2 = 0.10, p < 0.0001), likely
attributed to the higher relative contribution of new organic matter production and inputs post-
abandonment/conversion compared to the lower significance of such additions in landscapes
with greater NPP (i.e., in more temperate zones). In other words, the act of ALA/conversion
can have a much more dramatic impact on Mediterranean agroecosystems in terms of SOM
than in other regions. In accordance with this, there is a much lower relative rate of SOC stock
increase in Continental soils (0.27% yr–1, n = 165, R2 = 0.02, p = n.s.), and even a negative rate
(i.e., steady SOC loss) in soils within Alpine regions (-0.13% yr–1, n = 80, R2 = 0.01, p = n.s.),
as the site conditions differ greatly from Mediterranean conditions. However, these trends are
highly variable and non-significant at p < 0.05, especially in absolute terms. The only region
with greater relative rates of SOC change than the Mediterranean was the Pannonian region
(1.76% yr–1, n = 40, R2 = 0.38, p < 0.0001), where warm, wet, dry, and cold fronts from the
Mediterranean, Alps, and Carpathians all converge in the sheltered basin. It is also worth noting
that the Mediterranean relative SCS rate reported here is much smaller than the rate identified
in CHAPTER III as a result of the inclusion of abandoned pastures, the data standardization to
full 0–30 cm profiles, and the quantification and analysis of SOC stock instead of
concentration.

4.4.2.2 Land use and management factors on SOC dynamics


Aside from abiotic factors across Europe influencing SOC response to ALA, such as climate
and topography, human driven factors also play a critical role especially at smaller spatial
scales (i.e., the plot or landscape). In this study, I explored the influence of past land use
classification (whether cropland or pasture), post-abandonment/conversion land management
regimes (whether natural succession or assisted restoration), and past crop type (whether
woody or annual) (Figure 27). Sites that were croplands (n = 393) before
abandonment/conversion had a notably greater rate of SOC increase over time relatively
(1.52% yr–1, n = 393, R2 = 0.27, p < 0.0001) and absolutely (0.38 Mg ha–1 yr–1, n = 393, R2 =
0.18, p < 0.0001) than sites that were previously used as pastures (n = 145, p = n.s.). Pastures
are expected to have greater initial SOC stocks than croplands at the time of
abandonment/conversion, resulting in a lower or negative relative changes in SOC stock as
indicated in Figure 27.d. Croplands, on the other hand, may receive more intensive agricultural
practices than pastures, including significant biomass removal and regular tillage, which
depletes SOC stocks and allows for greater positive relative changes in SOC stock following

128
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

the cessation of these practices (i.e., abandonment/conversion) (García et al., 2007). Although
there is variation among cropland types, with cereal cultivation receiving SOM friendly
management practices akin to pastures (e.g., manure application, stubble grazing, seed
fallowing) compared to woody croplands that receive poor SOM management practices (e.g.,
pruned branch losses), croplands as a whole are generally under more SOM degrading
management systems than pastures and meadows (Navas et al., 2012; Ruecker et al., 1998).
Pasture and grassland plants also have a longer growing season and are more efficient at
allocating carbon to soil through their roots than crops, both of which helps them overcome
their lower efficiency in converting CO2 into organic matter and further explains their greater
SOC stock (Kuzyakov & Domanski, 2000).

129
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

Figure 27. Absolute (left, Mg ha–1) and relative (right, %) change in SOC stock over time since
abandonment/conversion (yr) and their linear regressions for post-agricultural sites that:
underwent natural succession or assisted restoration after abandonment/conversion (a,b);
were either croplands or pastures before abandonment/conversion (c,d); had woody or annual
crops (if previously used as croplands) (e,f). Shaded areas represent respective 95%
confidence intervals. Numbers in parenthesis indicate sample sizes.

130
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

Similar to past land use, the influence of post-abandonment/conversion management systems


also produced divergent SOC responses in our dataset (Figure 27.a.b). Sites that were
abandoned from agriculture and left to undergo spontaneous ecological succession (n = 326)
exhibited a greater rate of change in SOC stock relatively (1.59% yr–1, R2 = 0.29, p < 0.0001)
and absolutely (0.35 Mg ha–1 yr–1, R2 = 0.14, p < 0.0001) compared to sites that were actively
restored or converted to new vegetation land covers (n = 194, p = n.s.), for example through
tree planting practices. The potential of each management approach for SOC accrual certainly
depends on conditions at the site of abandonment/conversion. For example, the long land use
history involving intensive agricultural in the Mediterranean biogeographical region often
requires specific forms of active restoration to overcome stalled vegetation recovery that
natural succession may lead to (Garcia-Franco et al., 2014; Ruiz-Navarro et al., 2009; Segura
et al., 2020, 2016). While past woody croplands exhibited an overall higher increase in SOC
across the whole dataset than past annual croplands (25.88±4.5 and 9.31±3.11 Mg ha–1,
respectively) (Figure 22), the SCS rate of both past crop types are similar (Figure 27.e.f). In
absolute terms, past woody (n = 126) and annual (n = 385) croplands sequestered SOC at rates
of 0.40 Mg ha–1 yr–1 (R2 = 0.40, p < 0.0001) and 0.39 Mg ha–1 yr–1 (R2 = 0.19, p < 0.0001)
respectively, while in relative terms the respective rates were 0.67% yr–1 (R2 = 0.05, p = 0.0143)
and 0.79% yr–1 (R2 = 0.07, p < 0.0001).

The results synthesized in this study represent one of the most comprehensive assessments of
the positive or negative impacts of ALA on SCS at a continental scale. However, due to the
diversity in the original research aims of all the studies providing data and the methodologies
used here, there remains uncertainty in many of the sequestration rates estimated. For example,
uneven distribution in time since abandonment/conversion, geographic location (site
coordinates), and human management factors may skew the overall trends towards more
positive or negative rates. This is in addition to the fact that ALA is already biased towards
more degraded soils that may be less conducive for SCS than highly productive agricultural
soils that have the potential for positive legacy fertilizer effects if agricultural practices cease
and the landscape is restored. Furthermore, because many of the collected studies did not
provide important variables like bulk density, coarse material content, SOC concentration at
each depth examined, the necessary standardization steps used here have undoubtably reduced
accuracy. And lastly, neglect of subsoil samples in many studies and the standardization to the
0–30 cm depth for the entire dataset leaves subsoil SOC stocks unquantified during ALA. The
trends observed in topsoils cannot be assumed to hold true in subsoils. Further efforts should

131
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

focus on improving the dataset representativeness for all predictor factors and sample variables
(e.g., bulk density), and incorporating deeper SOC stocks for total soil carbon assessments in
Europe, where many soils reach far beyond 30 cm.

4.5 Conclusions
The widespread historical and ongoing agricultural land abandonment found across Europe has
resulted in slow, but steady increases in soil organic carbon stocks at an overall rate of 1.28%
yr–1, sequestering 0.32 Mg C ha–1 yr–1. However, large variabilities in rates are apparent, with
some post-agricultural landscapes losing SOC stock over time. In general, sites with low initial
stock had greater potential for SOC accumulation while sites with high initial stock are
presumably closer to SOC saturation and unlikely to exhibit large relative increases post-
abandonment/conversion. Climatic conditions and biogeographical regions influence the
likelihood of an abandoned/converted agro-landscape to accumulate SOC, with specific
combinations likely driving the observed divergent SOC stock accumulation/loss rates. Past
land use (cropland vs. pasture) and post-abandonment/conversion land management strategy
employed (natural vs. assisted) also produced divergent responses in SOC change, implying
that croplands managed through natural succession would show the greatest SOC accrual while
pastures that are actively converted (e.g., afforestation) would result in the lowest increases in
SOC, or even losses. The high variability and divergencies in post-abandonment/conversion
SOC dynamics must be considered in sustainable land use planning that strives to incorporate
the ecological and climate change mitigation benefits of agricultural land abandonment, taking
into account site-specific conditions and past and present land management histories to avoid
detrimental impacts for soil health and lost opportunities for ecosystem restoration.

132
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

4.6 References
1. Alberti, G., Leronni, V., Piazzi, M., Petrella, F., Mairota, P., Peressotti, A., Piussi, P., Valentini, R., Gristina,
L., Mantia, T.L., Novara, A., Rühl, J., 2011. Impact of woody encroachment on soil organic carbon and
nitrogen in abandoned agricultural lands along a rainfall gradient in Italy. Regional Environmental Change.
https://doi.org/10.1007/s10113-011-0229-6

2. Amelung, W., Bossio, D., de Vries, W., Kögel-Knabner, I., Lehmann, J., Amundson, R., Bol, R., Collins, C.,
Lal, R., Leifeld, J., Minasny, B., Pan, G., Paustian, K., Rumpel, C., Sanderman, J., van Groenigen, J.W.,
Mooney, S., van Wesemael, B., Wander, M., Chabbi, A., 2020. Towards a global-scale soil climate mitigation
strategy. Nature Communications 11, 5427. https://doi.org/10.1038/s41467-020-18887-7

3. Bárcena, T.G., Kiær, L.P., Vesterdal, L., Stefánsdóttir, H.M., Gundersen, P., Sigurdsson, B.D., 2014. Soil
carbon stock change following afforestation in Northern Europe: A meta-analysis. Global Change Biology.
https://doi.org/10.1111/gcb.12576

4. Beilin, R., Lindborg, R., Stenseke, M., Pereira, H.M., Llausàs, A., Slätmo, E., Cerqueira, Y., Navarro, L.,
Rodrigues, P., Reichelt, N., Munro, N., Queiroz, C., 2014. Analysing how drivers of agricultural land
abandonment affect biodiversity and cultural landscapes using case studies from Scandinavia, Iberia and
Oceania. Land Use Policy 36, 60–72. https://doi.org/10.1016/J.LANDUSEPOL.2013.07.003

5. Benjamin, K., Bouchard, A., Domon, G., 2007. Abandoned farmlands as components of rural landscapes: An
analysis of perceptions and representations. Landscape and Urban Planning 83, 228–244.
https://doi.org/10.1016/j.landurbplan.2007.04.009

6. Bonet, A., 2004. Secondary succession of semi-arid Mediterranean old-fields in south-eastern Spain: insights
for conservation and restoration of degraded lands. Journal of Arid Environments 56, 213–233.
https://doi.org/10.1016/S0140-1963(03)00048-X

7. Bossio, D.A., Cook-Patton, S.C., Ellis, P.W., Fargione, J., Sanderman, J., Smith, P., Wood, S., Zomer, R.J.,
von Unger, M., Emmer, I.M., Griscom, B.W., 2020. The role of soil carbon in natural climate solutions.
Nature Sustainability. https://doi.org/10.1038/s41893-020-0491-z

8. Bradford, M.A., Carey, C.J., Atwood, L., Bossio, D., Fenichel, E.P., Gennet, S., Fargione, J., Fisher, J.R.B.,
Fuller, E., Kane, D.A., Lehmann, J., Oldfield, E.E., Ordway, E.M., Rudek, J., Sanderman, J., Wood, S.A.,
2019. Soil carbon science for policy and practice. Nature Sustainability. https://doi.org/10.1038/s41893-019-
0431-y

9. Breuer, L., Huisman, J.A., Keller, T., Frede, H.-G., 2006. Impact of a conversion from cropland to grassland
on C and N storage and related soil properties: Analysis of a 60-year chronosequence. Geoderma 133, 6–18.
https://doi.org/10.1016/J.GEODERMA.2006.03.033

10. Calcagno, V., Mazancourt, C. de, 2010. “glmulti”: An R Package for Easy Automated Model Selection with
(Generalized) Linear Models. Journal of Statistical Software 34. https://doi.org/10.18637/jss.v034.i12

11. Campbell, J.E., Lobell, D.B., Genova, R.C., Field, C.B., 2008. The global potential of bioenergy on
abandoned agriculture lands. Environmental Science and Technology 42, 5791–5794.
https://doi.org/10.1021/es800052w

12. Carlson, K.M., Gerber, J.S., Mueller, N.D., Herrero, M., MacDonald, G.K., Brauman, K.A., Havlik, P.,
O’Connell, C.S., Johnson, J.A., Saatchi, S., West, P.C., 2017. Greenhouse gas emissions intensity of global
croplands. Nature Climate Change 7, 63–68. https://doi.org/10.1038/nclimate3158

13. Conant, R.T., Paustian, K., Elliott, E.T., 2001. GRASSLAND MANAGEMENT AND CONVERSION INTO
GRASSLAND: EFFECTS ON SOIL CARBON, Ecological Applications.

14. Cook-Patton, S.C., Leavitt, S.M., Gibbs, D., Harris, N.L., Lister, K., Anderson-Teixeira, K.J., Briggs, R.D.,
Chazdon, R.L., Crowther, T.W., Ellis, P.W., Griscom, H.P., Herrmann, V., Holl, K.D., Houghton, R.A.,
Larrosa, C., Lomax, G., Lucas, R., Madsen, P., Malhi, Y., Paquette, A., Parker, J.D., Paul, K., Routh, D.,
Roxburgh, S., Saatchi, S., van den Hoogen, J., Walker, W.S., Wheeler, C.E., Wood, S.A., Xu, L., Griscom,

133
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

B.W., 2020. Mapping carbon accumulation potential from global natural forest regrowth. Nature 585, 545–
550. https://doi.org/10.1038/s41586-020-2686-x

15. Cramer, V.A., Hobbs, R.J., Standish, R.J., 2008. What’s new about old fields? Land abandonment and
ecosystem assembly. Trends in Ecology & Evolution 23, 104–112.
https://doi.org/10.1016/J.TREE.2007.10.005

16. Deng, L., Liu, G. bin, Shangguan, Z. ping, 2014. Land-use conversion and changing soil carbon stocks in
China’s “Grain-for-Green” Program: A synthesis. Global Change Biology. https://doi.org/10.1111/gcb.12508

17. Deng, L., Zhu, G., Tang, Z., Shangguan, Z., 2016. Global patterns of the effects of land-use changes on soil
carbon stocks. Global Ecology and Conservation 5, 127–138. https://doi.org/10.1016/J.GECCO.2015.12.004

18. EC, 2012. Report from the Commission to the European Parliament, the Council, the European Economic
and Social Committee and the Committee of the Regions, The implementation of the Soil Thematic Strategy
and on going activities. (COM(2012) 46 final) Official Journal of the European Union, Brussels.

19. EEA, 2016. Biogeographical regions [WWW Document]. URL https://www.eea.europa.eu/data-and-


maps/data/biogeographical-regions-europe-3#tab-metadata (accessed 7.4.21).

20. Estel, S., Kuemmerle, T., Alcántara, C., Levers, C., Prishchepov, A., Hostert, P., 2015. Mapping farmland
abandonment and recultivation across Europe using MODIS NDVI time series. Remote Sensing of
Environment 163, 312–325. https://doi.org/10.1016/j.rse.2015.03.028

21. EU NIR, 2021. Annual European Union greenhouse gas inventory 1990–2019 and inventory report 2021
(National inventory reports (NIR)). European Commission, DG Climate Action European Environment
Agency.

22. Gabarrón-Galeote, M.A., Trigalet, S., Wesemael, B. van, 2015. Soil organic carbon evolution after land
abandonment along a precipitation gradient in southern Spain. Agriculture, Ecosystems & Environment 199,
114–123. https://doi.org/10.1016/J.AGEE.2014.08.027

23. García, H., Tarrasón, D., Mayol, M., Male-Bascompte, N., Riba, M., 2007. Patterns of variability in soil
properties and vegetation cover following abandonment of olive groves in Catalonia (NE Spain). Acta
Oecologica 31, 316–324. https://doi.org/10.1016/J.ACTAO.2007.01.001

24. Garcia-Franco, N., Wiesmeier, M., Goberna, M., Martínez-Mena, M., Albaladejo, J., 2014. Carbon dynamics
after afforestation of semiarid shrublands: Implications of site preparation techniques. Forest Ecology and
Management 319, 107–115. https://doi.org/10.1016/j.foreco.2014.01.043

25. Gardi, C., Visioli, G., Conti, F. D., Scotti, M., Menta, C., & Bodini, A. (2016). High Nature Value Farmland:
Assessment of Soil Organic Carbon in Europe. Frontiers in Environmental Science, 4.
https://doi.org/10.3389/fenvs.2016.00047

26. Guidi, C., Magid, J., Rodeghiero, M., Gianelle, D., Vesterdal, L., 2014. Effects of forest expansion on
mountain grassland: changes within soil organic carbon fractions. Plant and Soil 385, 373–387.
https://doi.org/10.1007/s11104-014-2315-2

27. Guo, L.B., Gifford, R.M., 2002. Soil carbon stocks and land use change: A meta analysis. Global Change
Biology. https://doi.org/10.1046/j.1354-1013.2002.00486.x

28. Heinimann, A., Mertz, O., Frolking, S., Christensen, A.E., Hurni, K., Sedano, F., Chini, L.P., Sahajpal, R.,
Hansen, M., Hurtt, G., 2017. A global view of shifting cultivation: Recent, current, and future extent. PLoS
ONE 12. https://doi.org/10.1371/journal.pone.0184479

29. Hoogmoed, M., Cunningham, S.C., Thomson, J.R., Baker, P.J., Beringer, J., Cavagnaro, T.R., 2012. Does
afforestation of pastures increase sequestration of soil carbon in Mediterranean climates? Agriculture,
Ecosystems & Environment 159, 176–183. https://doi.org/10.1016/J.AGEE.2012.07.011

134
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

30. Hu, P., Liu, S., Ye, Y., Zhang, W., He, X., Su, Y., Wang, K., 2018. Soil carbon and nitrogen accumulation
following agricultural abandonment in a subtropical karst region. Applied Soil Ecology.
https://doi.org/10.1016/j.apsoil.2018.09.003

31. Ibáñez, J.J., Zinck, J.A., Dazzi, C., 2013. Soil geography and diversity of the European biogeographical
regions. Geoderma 192, 142–153. https://doi.org/10.1016/j.geoderma.2012.07.024

32. Jackson, R.B., Banner, J.L., Jobbágy, E.G., Pockman, W.T., Wall, D.H., 2002. Ecosystem carbon loss with
woody plant invasion of grasslands. Nature 418, 623–626. https://doi.org/10.1038/nature00910

33. Jobbágy, E., Jackson, R., 2000. THE VERTICAL DISTRIBUTION OF SOIL ORGANIC CARBON AND
ITS RELATION TO CLIMATE AND VEGETATION. Ecological Applications 10, 423–436.

34. Kämpf, I., Hölzel, N., Störrle, M., Broll, G., Kiehl, K., 2016. Potential of temperate agricultural soils for
carbon sequestration: A meta-analysis of land-use effects. Science of The Total Environment 566–567, 428–
435. https://doi.org/10.1016/J.SCITOTENV.2016.05.067

35. Koch, A., Brierley, C., Maslin, M.M., Lewis, S.L., 2019. Earth system impacts of the European arrival and
Great Dying in the Americas after 1492. Quaternary Science Reviews 207, 13–36.
https://doi.org/10.1016/J.QUASCIREV.2018.12.004

36. Kuemmerle, T., Olofsson, P., Chaskovskyy, O., Baumann, M., Ostapowicz, K., Woodcock, C.E., Houghton,
R.A., Hostert, P., Keeton, W.S., Radeloff, V.C., 2011. Post-Soviet farmland abandonment, forest recovery,
and carbon sequestration in western Ukraine. Global Change Biology. https://doi.org/10.1111/j.1365-
2486.2010.02333.x

37. Kuzyakov, Y., & Domanski, G. (2000). Carbon input by plants into the soil. Review. Journal of Plant
Nutrition and Soil Science, 163(4), 421–431. https://doi.org/10.1002/1522-2624(200008)163:4<421::AID-
JPLN421>3.0.CO;2-R

38. La Mantia, T., Gristina, L., Rivaldo, E., Pasta, S., Novara, A., Rühl, J., 2013. The effects of post-pasture
woody plant colonization on soil and aboveground litter carbon and nitrogen along a bioclimatic transect.
IForest 6, 238–246. https://doi.org/10.3832/ifor0811-006

39. Laganière, J., Angers, D.A., Paré, D., 2010. Carbon accumulation in agricultural soils after afforestation: A
meta-analysis. Global Change Biology 16, 439–453. https://doi.org/10.1111/j.1365-2486.2009.01930.x

40. Lal, R., 2013. Intensive Agriculture and the Soil Carbon Pool. Journal of Crop Improvement 27, 735–751.
https://doi.org/10.1080/15427528.2013.845053

41. Lal, R., 2004. Soil Carbon Sequestration Impacts on Global Climate Change and Food Security. Science
(1979) 304, 1623–1627. https://doi.org/10.1126/science.1097396

42. Lasanta, T., Arnáez, J., Pascual, N., Ruiz-Flaño, P., Errea, M.P., Lana-Renault, N., 2017. Space–time process
and drivers of land abandonment in Europe. CATENA 149, 810–823.
https://doi.org/10.1016/J.CATENA.2016.02.024

43. Lasanta, T., Nadal-Romero, E., Arnáez, J., 2015. Managing abandoned farmland to control the impact of re-
vegetation on the environment. The state of the art in Europe. Environmental Science & Policy 52, 99–109.
https://doi.org/10.1016/J.ENVSCI.2015.05.012

44. Lavalle, C., Silva, F.B.E., Baranzelli, C., Jacobs-Crisioni, C., Kompil, M., Perpiña Castillo, C., Vizcaino, P.,
Ribeiro Barranco, R., Vandecasteele, I., Kavalov, B., Aurambout, J.-P., Kucas, A., Siragusa, A., Auteri, D.,
2020. The LUISA Territorial Modelling Platform and Urban Data Platform: An EU-Wide Holistic Approach,
in: Medeiros, E. (Ed.), Territorial Impact Assessment. Advances in Spatial Science (The Regional Science
Series). Springer, Cham. https://doi.org/10.1007/978-3-030-54502-4_10

45. Levers, C., Schneider, M., Prishchepov, A. V., Estel, S., Kuemmerle, T., 2018. Spatial variation in
determinants of agricultural land abandonment in Europe. Science of the Total Environment.
https://doi.org/10.1016/j.scitotenv.2018.06.326

135
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

46. Li, D., Niu, S., Luo, Y., 2012. Global patterns of the dynamics of soil carbon and nitrogen stocks following
afforestation: A meta-analysis. New Phytologist 195, 172–181. https://doi.org/10.1111/j.1469-
8137.2012.04150.x

47. Li, S., Li, X., 2017. Global understanding of farmland abandonment: A review and prospects. Journal of
Geographical Sciences. https://doi.org/10.1007/s11442-017-1426-0

48. Li, W., Ciais, P., Guenet, B., Peng, S., Chang, J., Chaplot, V., Khudyaev, S., Peregon, A., Piao, S., Wang, Y.,
Yue, C., 2018. Temporal response of soil organic carbon after grassland‐related land‐use change. Global
Change Biology. https://doi.org/10.1111/gcb.14328

49. Manrique, L.A., Jones, C.A., 1991. Bulk Density of Soils in Relation to Soil Physical and Chemical
Properties. Soil Science Society of America Journal 55.
https://doi.org/10.2136/sssaj1991.03615995005500020030x

50. Martinez-Duro, E., Ferrandis, P., Escudero, A., Luzuriaga, A.L., Herranz, J.M., 2010. Secondary old-field
succession in an ecosystem with restrictive soils: Does time from abandonment matter? Applied Vegetation
Science. https://doi.org/10.1111/j.1654-109X.2009.01064.x

51. Montanarella, L., Panagos, P., 2021. The relevance of sustainable soil management within the European
Green Deal. Land Use Policy 100. https://doi.org/10.1016/j.landusepol.2020.104950

52. Moreno, J.L., Torres, I.F., García, C., López-Mondéjar, R., Bastida, F., 2019. Land use shapes the resistance
of the soil microbial community and the C cycling response to drought in a semi-arid area. Science of The
Total Environment 648, 1018–1030. https://doi.org/10.1016/J.SCITOTENV.2018.08.214

53. Morrone, J.J., 2018. The spectre of biogeographical regionalization. Journal of Biogeography 45.
https://doi.org/10.1111/jbi.13135

54. Nadal-Romero, E., Cammeraat, E., Pérez-Cardiel, E., Lasanta, T., 2016. How do soil organic carbon stocks
change after cropland abandonment in Mediterranean humid mountain areas? Science of The Total
Environment 566–567, 741–752. https://doi.org/10.1016/J.SCITOTENV.2016.05.031

55. Navarro, L.M., Pereira, H.M., 2012. Rewilding abandoned landscapes in Europe, in: Rewilding European
Landscapes. Springer International Publishing, pp. 3–23. https://doi.org/10.1007/978-3-319-12039-3_1

56. Navas, A., Gaspar, L., Quijano, L., López-Vicente, M., Machín, J., 2012. Patterns of soil organic carbon and
nitrogen in relation to soil movement under different land uses in mountain fields (South Central Pyrenees).
Catena (Amst) 94, 43–52. https://doi.org/10.1016/j.catena.2011.05.012

57. Palmero-Iniesta, M., Pino, J., Pesquer, L., & Espelta, J. M. (2021). Recent forest area increase in Europe:
expanding and regenerating forests differ in their regional patterns, drivers and productivity trends. European
Journal of Forest Research, 140(4), 793–805. https://doi.org/10.1007/s10342-021-01366-z

58. Paul, K.I., Polglase, P.J., Nyakuengama, J.G., Khanna, P.K., 2002. Change in soil carbon following
afforestation. Forest Ecology and Management 168, 241–257. https://doi.org/10.1016/S0378-
1127(01)00740-X

59. Pellis, G., Chiti, T., Rey, A., Curiel Yuste, J., Trotta, C., Papale, D., 2019. The ecosystem carbon sink
implications of mountain forest expansion into abandoned grazing land: The role of subsoil and climatic
factors. Science of The Total Environment 672, 106–120.
https://doi.org/10.1016/J.SCITOTENV.2019.03.329

60. Perpiña Castillo, C., Jacobs-Crisioni, C., Diogo, V., Lavalle, C., 2021. Modelling agricultural land
abandonment in a fine spatial resolution multi-level land-use model: An application for the EU.
Environmental Modelling and Software 136. https://doi.org/10.1016/j.envsoft.2020.104946

61. Plieninger, T., Hui, C., Gaertner, M., Huntsinger, L., 2014. The impact of land abandonment on species
richness and abundance in the Mediterranean Basin: A meta-analysis. PLoS ONE.
https://doi.org/10.1371/journal.pone.0098355

136
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

62. Poeplau, C., Don, A., Vesterdal, L., Leifeld, J., van Wesemael, B., Schumacher, J., Gensior, A., 2011.
Temporal dynamics of soil organic carbon after land-use change in the temperate zone - carbon response
functions as a model approach. Global Change Biology. https://doi.org/10.1111/j.1365-2486.2011.02408.x

63. Post, W.M., Kwon, K.C., 2000. Soil carbon sequestration and land‐use change: processes and potential.
Global Change Biology 6. https://doi.org/10.1046/j.1365-2486.2000.00308.x

64. Pribyl, D.W., 2010. A critical review of the conventional SOC to SOM conversion factor. Geoderma 156,
75–83. https://doi.org/10.1016/j.geoderma.2010.02.003

65. Queiroz, C., Beilin, R., Folke, C., Lindborg, R., 2014. Farmland abandonment: Threat or opportunity for
biodiversity conservation? A global review. Frontiers in Ecology and the Environment.
https://doi.org/10.1890/120348

66. R Core Team, 2020. R: A language and environment for statistical computing. R Foundation for Statistical
Computing.

67. Robledano-Aymerich, F., Romero-Díaz, A., Belmonte-Serrato, F., Zapata-Pérez, V.M., Martínez-Hernández,
C., Martínez-López, V., 2014. Ecogeomorphological consequences of land abandonment in semiarid
Mediterranean areas: Integrated assessment of physical evolution and biodiversity. Agriculture, Ecosystems
& Environment 197, 222–242. https://doi.org/10.1016/J.AGEE.2014.08.006

68. Rodrigues, L., Fohrafellner, J., Hardy, B., Huyghebaert, B., & Leifeld, J., 2021. Towards climate-smart
sustainable management of agricultural soils. Deliverable 2.3 Synthesis on estimates of achievable soil carbon
sequestration on agricutural land across Europe.

69. Ruecker, G., Schad, P., Alcubilla, M.M., Ferrer, C., 1998. Natural regeneration of degraded soils and site
changes on abandoned agricultural terraces in Mediterranean Spain. Land Degradation & Development 9,
179–188.

70. Ruiz-Navarro, A., Barberá, G.G., Navarro-Cano, J.A., Albaladejo, J., Castillo, V.M., 2009. Soil dynamics in
Pinus halepensis reforestation: Effect of microenvironments and previous land use. Geoderma 153, 353–361.
https://doi.org/10.1016/j.geoderma.2009.08.024

71. Schierhorn, F., Müller, D., Beringer, T., Prishchepov, A. V., Kuemmerle, T., Balmann, A., 2013. Post-Soviet
cropland abandonment and carbon sequestration in European Russia, Ukraine, and Belarus. Global
Biogeochemical Cycles. https://doi.org/10.1002/2013GB004654

72. Schröder, P., Beckers, B., Daniels, S., Gnädinger, F., Maestri, E., Marmiroli, N., Mench, M., Millan, R.,
Obermeier, M.M., Oustriere, N., Persson, T., Poschenrieder, C., Rineau, F., Rutkowska, B., Schmid, T.,
Szulc, W., Witters, N., Sæbø, A., 2018. Intensify production, transform biomass to energy and novel goods
and protect soils in Europe—A vision how to mobilize marginal lands. Science of The Total Environment
616–617, 1101–1123. https://doi.org/10.1016/J.SCITOTENV.2017.10.209

73. Segura, C., Jiménez, M.N., Nieto, O., Navarro, F.B., Fernández-Ondoño, E., 2016. Changes in soil organic
carbon over 20 years after afforestation in semiarid SE Spain. Forest Ecology and Management.
https://doi.org/10.1016/j.foreco.2016.09.035

74. Segura, C., Navarro, F.B., Jiménez, M.N., Fernández-Ondoño, E., 2020. Implications of afforestation vs.
secondary succession for soil properties under a semiarid climate. Science of the Total Environment 704.
https://doi.org/10.1016/j.scitotenv.2019.135393

75. Shi, S., Zhang, W., Zhang, P., Yu, Y., Ding, F., 2013. A synthesis of change in deep soil organic carbon stores
with afforestation of agricultural soils. Forest Ecology and Management 296, 53–63.
https://doi.org/10.1016/J.FORECO.2013.01.026

76. Silver, W.L., Ostertag, R., Lugo, A.E., 2000. The Potential for Carbon Sequestration Through Reforestation
of Abandoned Tropical Agricultural and Pasture Lands.

77. Stewart, C.E., Paustian, K., Conant, R.T., Plante, A.F., Six, J., 2007. Soil carbon saturation: Concept,
evidence and evaluation. Biogeochemistry 86, 19–31. https://doi.org/10.1007/s10533-007-9140-0

137
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

78. Sulman, B.N., Harden, J., He, Y., Treat, C., Koven, C., Mishra, U., O’Donnell, J.A., Nave, L.E., 2020. Land
Use and Land Cover Affect the Depth Distribution of Soil Carbon: Insights From a Large Database of Soil
Profiles. Frontiers in Environmental Science 8. https://doi.org/10.3389/fenvs.2020.00146

79. Ustaoglu, E., Collier, M.J., 2018. Farmland abandonment in Europe: an overview of drivers, consequences,
and assessment of the sustainability implications. Environmental Reviews. https://doi.org/10.1139/er-2018-
0001

80. Vermeulen, S., Bossio, D., Lehmann, J., Luu, P., Paustian, K., Webb, C., Augé, F., Bacudo, I., Baedeker, T.,
Havemann, T., Jones, C., King, R., Reddy, M., Sunga, I., Von Unger, M., Warnken, M., 2019. A global
agenda for collective action on soil carbon. Nature Sustainability. https://doi.org/10.1038/s41893-018-0212-
z

81. Walker, L.R., Wardle, D.A., Bardgett, R.D., Clarkson, B.D., 2010. The use of chronosequences in studies of
ecological succession and soil development. Journal of Ecology. https://doi.org/10.1111/j.1365-
2745.2010.01664.x

82. Wertebach, T.M., Hölzel, N., Kämpf, I., Yurtaev, A., Tupitsin, S., Kiehl, K., Kamp, J., Kleinebecker, T.,
2017. Soil carbon sequestration due to post-Soviet cropland abandonment: estimates from a large-scale soil
organic carbon field inventory. Global Change Biology 23, 3729–3741. https://doi.org/10.1111/gcb.13650

83. Xie, Z., Game, E.T., Hobbs, R.J., Pannell, D.J., Phinn, S.R., McDonald-Madden, E., 2020. Conservation
opportunities on uncontested lands. Nature Sustainability 3, 9–15. https://doi.org/10.1038/s41893-019-0433-
9

84. Yang, Y., Luo, Y., Finzi, A.C., 2011. Carbon and nitrogen dynamics during forest stand development: a
global synthesis. New Phytologist 190, 977–989. https://doi.org/10.1111/j.1469-8137.2011.03645.x

85. Yin, H., Brandão, A., Buchner, J., Helmers, D., Iuliano, B.G., Kimambo, N.E., Lewińska, K.E., Razenkova,
E., Rizayeva, A., Rogova, N., Spawn, S.A., Xie, Y., Radeloff, V.C., 2020. Monitoring cropland abandonment
with Landsat time series. Remote Sensing of Environment 246, 111873.
https://doi.org/10.1016/j.rse.2020.111873

138
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

4.7 Supplementary materials


Table 10. Results from Kruskal–Wallis tests of significance for factor effects on SOC stock
change.

Factor df Chi-square p-value


CTRL 379.95 240 2.144e-08
AGE 166.5 95 8.101e-06
PLU 34.657 1 3.932e-09
MGMT 10.678 3 0.0136
CROP 65.589 2 5.721e-15
BIOGEO 91.092 5 < 2.2e-16

Figure 28. Model-averaged importance of terms following multi-model analysis (generalized


linear model). Information criterion set to AICc, and search set to “exhaustive”. Factors
exceeding the red line have the most importance among all models. SCS, soil carbon
sequestration; MGMT, management regime; PLU, past land use; CROP, past crop type;
CTRL, initial SOC stock; BIOGEO, biogeographical region; AGE, time since
abandonment/conversion.

139
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

Table 11. Summary of the best-fitting linear model following multi-model analysis. 43 models
were needed to reach 95% of evidence weight, and convergence achieved after 540
generations. SCS, soil carbon sequestration; MGMT, management regime; PLU, past land
use; CROP, past crop type; CTRL, initial SOC stock; BIOGEO, biogeographical region; AGE,
time since abandonment/conversion.

140
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

Table 12. List of papers included in the European synthesis.

1 Alberti, G., Leronni, V., Piazzi, M., Petrella, F., Mairota, P., Peressotti, A., Piussi, P.,
Valentini, R., Gristina, L., Mantia, T.L., Novara, A., Rühl, J., 2011. Impact of woody
encroachment on soil organic carbon and nitrogen in abandoned agricultural lands along
a rainfall gradient in Italy. Reg. Environ. Chang. https://doi.org/10.1007/s10113-011-
0229-6
2 Alberti, G., Peressotti, A., Piussi, P. & Zerbi, G. Forest ecosystem carbon accumulation
during a secondary succession in the Eastern Prealps of Italy. Forestry 81, 1–11 (2008).
3 Alías, J.C., Mejías, J.A. and Chaves, N., 2022. Effect of Cropland Abandonment on Soil
Carbon Stock in an Agroforestry System in Southwestern Spain. Land, 11(3), p.425.
4 Apostolakis, A., Panakoulia, S., Nikolaidis, N.P., Paranychianakis, N. V., 2017. Shifts in
soil structure and soil organic matter in a chronosequence of set-aside fields. Soil Tillage
Res. 174, 113–119. https://doi.org/10.1016/J.STILL.2017.07.004
5 Ashwood, F., Watts, K., Park, K., Fuentes‐Montemayor, E., Benham, S. and
Vanguelova, E.I., 2019. Woodland restoration on agricultural land: long‐term impacts on
soil quality. Restoration Ecology, 27(6), pp.1381-1392.
6 Badalamenti, E., Battipaglia, G., Gristina, L., Novara, A., Ruhl, J., Sala, G., Sapienza, L.,
Valentini, R., La Mantia, T., 2019. Carbon stock increases up to old growth forest along
a secondary succession in Mediterranean island ecosystems. PLoS One 14.
https://doi.org/10.1371/journal.pone.0220194
7 Bárcena, T.G., Gundersen, P. and Vesterdal, L., 2014. Afforestation effects on SOC in
former cropland: oak and spruce chronosequences resampled after 13 years. Global
Change Biology, 20(9), pp.2938-2952.
8 Bell, S.M., Terrer, C., Barriocanal, C., Jackson, R.B., Rosell-Melé, A., 2021. Soil
organic carbon accumulation rates on Mediterranean abandoned agricultural lands. Sci.
Total Environ. 759. https://doi.org/10.1016/j.scitotenv.2020.143535
9 Biasi, R., Farina, R. and Brunori, E., 2021. Family Farming Plays an Essential Role in
Preserving Soil Functionality: A Study on Active Managed and Abandoned Traditional
Tree Crop-Based Systems. Sustainability, 13(7), p.3967.
10 Black, K., Byrne, K.A., Mencuccini, M., Tobin, B., Nieuwenhuis, M., Reidy, B., Bolger,
T., Saiz, G., Green, C., Farrell, E.T. and Osborne, B., 2009. Carbon stock and stock
changes across a Sitka spruce chronosequence on surface-water gley soils. Forestry,
82(3), pp.255-272.
11 Blanco-Moure, N., Gracia, R., Bielsa, A.C., López, M.V., 2016. Soil organic matter
fractions as affected by tillage and soil texture under semiarid Mediterranean conditions.
Soil Tillage Res. 155, 381–389. https://doi.org/10.1016/J.STILL.2015.08.011
12 Blanco-Moure, N., Moret-Fernández, D. and López, M.V., 2012. Dynamics of aggregate
destabilization by water in soils under long-term conservation tillage in semiarid Spain.
Catena, 99, pp.34-41.
13 Boecker, D., Centeri, C., Welp, G. and Möseler, B.M., 2015. Parallels of secondary
grassland succession and soil regeneration in a chronosequence of central-Hungarian old
fields. Folia Geobotanica, 50(2), pp.91-106.
14 Bonet, A., 2004. Secondary succession of semi-arid Mediterranean old-fields in south-
eastern Spain: insights for conservation and restoration of degraded lands. J. Arid
Environ. 56, 213–233. https://doi.org/10.1016/S0140-1963(03)00048-X
15 Breuer, L., Huisman, J.A., Keller, T. and Frede, H.G., 2006. Impact of a conversion from
cropland to grassland on C and N storage and related soil properties: Analysis of a 60-
year chronosequence. Geoderma, 133(1-2), pp.6-18.

141
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

16 Cammeraat, L.., Imeson, A.., 1999. The evolution and significance of soil–vegetation
patterns following land abandonment and fire in Spain. CATENA 37, 107–127.
https://doi.org/10.1016/S0341-8162(98)00072-1
17 Cerdà, A., Ackermann, O., Terol, E. and Rodrigo-Comino, J., 2019. Impact of farmland
abandonment on water resources and soil conservation in citrus plantations in eastern
Spain. Water, 11(4), p.824.
18 Chiti, T., Blasi, E., Pellis, G., Perugini, L., Chiriacò Maria, V., Valentini, R., 2018. Soil
organic carbon pool’s contribution to climate change mitigation on marginal land of a
Mediterranean montane area in Italy. J. Environ. Manage.
https://doi.org/10.1016/j.jenvman.2018.04.093
19 Company, J., Valiente, N., Fortesa, J., García-Comendador, J., Lucas-Borja, M.E.,
Ortega, R., Miralles, I. and Estrany, J., 2022. Secondary succession and parent material
drive soil bacterial community composition in terraced abandoned olive groves from a
Mediterranean hyper-humid mountainous area. Agriculture, Ecosystems & Environment,
332, p.107932.
20 Danise, T., Andriuzzi, W.S., Battipaglia, G., Certini, G., Guggenberger, G., Innangi, M.,
Mastrolonardo, G., Niccoli, F., Pelleri, F. and Fioretto, A., 2021. Mixed-species
plantation effects on soil biological and chemical quality and tree growth of a former
agricultural land. Forests, 12(7), p.842.
21 De Baets, S., Meersmans, J., Vanacker, V., Quine, T.A., Van Oost, K., 2013. Spatial
variability and change in soil organic carbon stocks in response to recovery following
land abandonment and erosion in mountainous drylands. Soil Use Manag.
https://doi.org/10.1111/sum.12017
22 De Baets, S., Van Oost, K., Baumann, K., Meersmans, J., Vanacker, V., Rumpel, C.,
2012. Lignin signature as a function of land abandonment and erosion in dry luvisols of
SE Spain. CATENA 93, 78–86. https://doi.org/10.1016/J.CATENA.2012.01.015
23 Del Galdo, I., Six, J., Peressotti, A., Cotrufo, M.F., 2003. Assessing the impact of land-
use change on soil C sequestration in agricultural soils by means of organic matter
fractionation and stable C isotopes. Glob. Chang. Biol. 9, 1204–1213.
24 Diquélou, S. and Rozé, F., 1999. Implantation du genêt à balais, précédent cultural et
dynamique du sol dans les friches (Bretagne, France). Comptes Rendus de l'Académie
des Sciences-Series III-Sciences de la Vie, 322(8), pp.705-715.
25 Djuma, H., Bruggeman, A., Zissimos, A., Christoforou, I., Eliades, M., Zoumides, C.,
2020. The effect of agricultural abandonment and mountain terrace degradation on soil
organic carbon in a Mediterranean landscape.
https://doi.org/10.1016/j.catena.2020.104741
26 Don, A., Scholten, T. and Schulze, E.D., 2009. Conversion of cropland into grassland:
Implications for soil organic‐carbon stocks in two soils with different texture. Journal of
Plant Nutrition and Soil Science, 172(1), pp.53-62.
27 Dunjó, G., Pardini, G., Gispert, M., 2003. Land use change effects on abandoned
terraced soils in a Mediterranean catchment, NE Spain. CATENA 52, 23–37.
https://doi.org/10.1016/S0341-8162(02)00148-0
28 Dutoit, T., Decaens, T. and Alard, D., 1997. Successional changes and diversity of soil
macrofaunal communities on chalk grasslands in Upper-Normandy (France). Acta
Oecologica, 18(2), pp.135-149.
29 Emran, M., Gispert, M., Pardini, G., 2012. Patterns of soil organic carbon, glomalin and
structural stability in abandoned Mediterranean terraced lands. Eur. J. Soil Sci.
https://doi.org/10.1111/j.1365-2389.2012.01493.x

142
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

30 Ferlan, M., Alberti, G., Eler, K., Batič, F., Peressotti, A., Miglietta, F., Zaldei, A.,
Simončič, P. and Vodnik, D., 2011. Comparing carbon fluxes between different stages of
secondary succession of a karst grassland. Agriculture, ecosystems & environment,
140(1-2), pp.199-207.
31 Fernández‐Calviño, D., Nóvoa‐Muñoz, J.C., López‐Periago, E. and Arias‐Estévez, M.,
2008. Changes in copper content and distribution in young, old and abandoned vineyard
acid soils due to land use changes. Land degradation & development, 19(2), pp.165-177.
32 Fino, E., Blasi, E., Perugini, L., Pellis, G., Valentini, R. and Chiti, T., 2020. Is Soil
Contributing to Climate Change Mitigation during Woody Encroachment? A Case Study
on the Italian Alps. Forests, 11(8), p.887.
33 Ford, H., Garbutt, A., Jones, D.L. and Jones, L., 2012. Impacts of grazing abandonment
on ecosystem service provision: coastal grassland as a model system. Agriculture,
ecosystems & environment, 162, pp.108-115.
34 Francaviglia, R., Benedetti, A., Doro, L., Madrau, S., Ledda, L., 2014. Influence of land
use on soil quality and stratification ratios under agro-silvo-pastoral Mediterranean
management systems. Agric. Ecosyst. Environ. 183, 86–92.
https://doi.org/10.1016/j.agee.2013.10.026
35 Gabarrón-Galeote, M.A., Trigalet, S., Wesemael, B. van, 2015. Soil organic carbon
evolution after land abandonment along a precipitation gradient in southern Spain. Agric.
Ecosyst. Environ. 199, 114–123. https://doi.org/10.1016/J.AGEE.2014.08.027
36 García, H., Tarrasón, D., Mayol, M., Male-Bascompte, N., Riba, M., 2007. Patterns of
variability in soil properties and vegetation cover following abandonment of olive groves
in Catalonia (NE Spain). Acta Oecologica 31, 316–324.
https://doi.org/10.1016/J.ACTAO.2007.01.001
37 Guggenberger, G., Christensen, B.T. and Zech, W., 1994. Land‐use effects on the
composition of organic matter in particle‐size separates of soil: I. Lignin and
carbohydrate signature. European journal of soil Science, 45(4), pp.449-458.
38 Guidi, C., Vesterdal, L., Gianelle, D. and Rodeghiero, M., 2014. Changes in soil organic
carbon and nitrogen following forest expansion on grassland in the Southern Alps. Forest
ecology and management, 328, pp.103-116.
39 Hamer, U., Makeschin, F., Stadler, J. & Klotz, S. Soil organic matter and microbial
community structure in set-aside and intensively managed arable soils in NE-Saxony,
Germany. Appl. soil Ecol. 40, 465–475 (2008).
40 Hewelke, E., 2019. Influence of abandoning agricultural land use on hydrophysical
properties of sandy soil. Water, 11(3), p.525.
41 Hiltbrunner, D., Zimmermann, S. and Hagedorn, F., 2013. Afforestation with Norway
spruce on a subalpine pasture alters carbon dynamics but only moderately affects soil
carbon storage. Biogeochemistry, 115(1), pp.251-266.
42 Hontoria, C., Velásquez, R., Benito, M., Almorox, J., Moliner, A., 2009. Bradford-
reactive soil proteins and aggregate stability under abandoned versus tilled olive groves
in a semi-arid calcisol. Soil Biol. Biochem. 41, 1583–1585.
https://doi.org/10.1016/J.SOILBIO.2009.04.025
43 Hunziker, M., Caviezel, C. and Kuhn, N.J., 2017. Shrub encroachment by green alder on
subalpine pastures: Changes in mineral soil organic carbon characteristics. Catena, 157,
pp.35-46.
44 Kosmas, C., Gerontidis, S., Marathianou, M., 2000. The effect of land use change on
soils and vegetation over various lithological formations on Lesvos (Greece). Catena 40,
51–68.

143
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

45 Kozak, M. and Pudełko, R., 2021. Impact assessment of the long-term fallowed land on
agricultural soils and the possibility of their return to agriculture. Agriculture, 11(2),
p.148.
46 La Mantia, T., Gristina, L., Rivaldo, E., Pasta, S., Novara, A. and Ruehl, J., 2013. The
effects of post-pasture woody plant colonization on soil and aboveground litter carbon
and nitrogen along a bioclimatic transect. iForest-Biogeosciences and Forestry, 6(5),
p.238.
47 Lara-Gómez, M.A., Navarro-Cerrillo, R.M., Ceacero, C.J., Ruiz-Goméz, F.J., Díaz-
Hernández, J.L. and Palacios Rodriguez, G., 2020. Use of Aerial Laser Scanning to
Assess the Effect on C Sequestration of Oak (Quercus ilex L. subsp. ballota [Desf.]
Samp-Q. suber L.) Afforestation on Agricultural Land. Geosciences, 10(2), p.41.
48 Leifeld, J. and Kögel-Knabner, I., 2005. Soil organic matter fractions as early indicators
for carbon stock changes under different land-use?. Geoderma, 124(1-2), pp.143-155.
49 Lesschen, J.P., Cammeraat, L.H., Kooijman, A.M., van Wesemael, B., 2008.
Development of spatial heterogeneity in vegetation and soil properties after land
abandonment in a semi-arid ecosystem. J. Arid Environ. 72, 2082–2092.
https://doi.org/10.1016/J.JARIDENV.2008.06.006
50 Lichner, L., Felde, V.J., Büdel, B., Leue, M., Gerke, H.H., Ellerbrock, R.H., Kollár, J.,
Rodný, M., Šurda, P., Fodor, N. and Sándor, R., 2018. Effect of vegetation and its
succession on water repellency in sandy soils. Ecohydrology, 11(6), p.e1991.
51 Martínez-Hernández, C., Rodrigo-Comino, J., Romero-Díaz, A., 2017. Impact of
lithology and soil properties on abandoned dryland terraces during the early stages of soil
erosion by water in south-east Spain. Hydrol. Process. 31, 3095–3109.
https://doi.org/10.1002/hyp.11251
52 Martinez-Mena, M., Lopez, J., Almagro, M., Boix-Fayos, C., Albaladejo, J., 2008. Effect
of water erosion and cultivation on the soil carbon stock in a semiarid area of South-East
Spain. Soil Tillage Res. 99, 119–129. https://doi.org/10.1016/J.STILL.2008.01.009
53 Marzaioli, R., d’Ascoli, R., De Pascale, R.A. and Rutigliano, F.A., 2010. Soil quality in a
Mediterranean area of Southern Italy as related to different land use types. Applied Soil
Ecology, 44(3), pp.205-212.
54 Meyer, S., Leifeld, J., Bahn, M. and Fuhrer, J., 2012. Free and protected soil organic
carbon dynamics respond differently to abandonment of mountain grassland.
Biogeosciences, 9(2), pp.853-865.
55 Mielnik, L., Hewelke, E., Weber, J., Oktaba, L., Jonczak, J. and Podlasiński, M., 2021.
Changes in the soil hydrophobicity and structure of humic substances in sandy soil taken
out of cultivation. Agriculture, Ecosystems & Environment, 319, p.107554.
56 Moreno, J.L., Torres, I.F., García, C., López-Mondéjar, R., Bastida, F., 2019. Land use
shapes the resistance of the soil microbial community and the C cycling response to
drought in a semi-arid area. Sci. Total Environ. 648, 1018–1030.
https://doi.org/10.1016/J.SCITOTENV.2018.08.214
57 Navas, A., Gaspar, L., Quijano, L., López-Vicente, M., Machín, J., 2012. Patterns of soil
organic carbon and nitrogen in relation to soil movement under different land uses in
mountain fields (South Central Pyrenees). Catena 94, 43–52.
https://doi.org/10.1016/j.catena.2011.05.012
58 Novák, T.J., Incze, J., Spohn, M., Glina, B. and Giani, L., 2014. Soil and vegetation
transformation in abandoned vineyards of the Tokaj Nagy-Hill, Hungary. Catena, 123,
pp.88-98.
59 Novara, A., Gristina, L., Kuzyakov, Y., Schillaci, C., Laudicina, V.A., La Mantia, T.,
2013. Turnover and availability of soil organic carbon under different Mediterranean

144
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

land-uses as estimated by 13C natural abundance. Eur. J. Soil Sci. 64, 466–475.
https://doi.org/10.1111/ejss.12038
60 Novara, A., Gristina, L., La Mantia, T., Rühl, J., 2013. Carbon dynamics of soil organic
matter in bulk soil and aggregate fraction during secondary succession in a
Mediterranean environment. Geoderma 193–194, 213–221.
https://doi.org/10.1016/J.GEODERMA.2012.08.036
61 Ortiz, C., Vázquez, E., Rubio, A., Benito, M., Schindlbacher, A., Jandl, R., Butterbach-
Bahl, K. and Díaz-Pinés, E., 2016. Soil organic matter dynamics after afforestation of
mountain grasslands in both a Mediterranean and a temperate climate. Biogeochemistry,
131(3), pp.267-280.
62 Palacios-Rodríguez, G., Quinto, L., Lara-Gómez, M.A., Pérez-Romero, J., Recio, J.M.,
Álvarez-Romero, M., Cachinero-Vivar, A.M., Hernández-Navarro, S. and Navarro-
Cerrillo, R.M., 2022. Carbon Sequestration in Carob (Ceratonia siliqua L.) Plantations
under the EU Afforestation Program in Southern Spain Using Low-Density Aerial Laser
Scanning (ALS) Data. Forests, 13(2), p.285.
63 Palese, A.M., Magno, R., Casacchia, T., Curci, M., Baronti, S., Miglietta, F., Crecchio,
C., Xiloyannis, C., Sofo, A., 2013. Chemical, biochemical, and microbiological
properties of soils from abandoned and extensively cultivated olive orchards. Sci. World
J. 2013. https://doi.org/10.1155/2013/496278
64 Pardini, G., Gispert, M., Dunjó, G., 2003. Runoff erosion and nutrient depletion in five
Mediterranean soils of NE Spain under different land use. Sci. Total Environ. 309, 213–
224. https://doi.org/10.1016/S0048-9697(03)00007-X
65 Pärtel, M. and Helm, A., 2007. Invasion of woody species into temperate grasslands:
relationship with abiotic and biotic soil resource heterogeneity. Journal of Vegetation
Science, 18(1), pp.63-70.
66 Peco, B., Navarro, E., Carmona, C.P., Medina, N.G. and Marques, M.J., 2017. Effects of
grazing abandonment on soil multifunctionality: The role of plant functional traits.
Agriculture, Ecosystems & Environment, 249, pp.215-225.
67 Peichl, M., Leava, N.A. and Kiely, G., 2012. Above-and belowground ecosystem
biomass, carbon and nitrogen allocation in recently afforested grassland and adjacent
intensively managed grassland. Plant and Soil, 350(1), pp.281-296.
68 Pellegrino, E., Bosco, S., Ciccolini, V., Pistocchi, C., Sabbatini, T., Silvestri, N., Bonari,
E., 2015. Agricultural abandonment in Mediterranean reclaimed peaty soils: long-term
effects on soil chemical properties, arbuscular mycorrhizas and CO2 flux. Agric.
Ecosyst. Environ. 199, 164–175. https://doi.org/10.1016/J.AGEE.2014.09.004
69 Pellis, G., Chiti, T., Rey, A., Yuste, J.C., Trotta, C. and Papale, D., 2019. The ecosystem
carbon sink implications of mountain forest expansion into abandoned grazing land: The
role of subsoil and climatic factors. Science of The Total Environment, 672, pp.106-120.
70 Plassart, P., Vinceslas, M.A., Gangneux, C., Mercier, A., Barray, S. and Laval, K., 2008.
Molecular and functional responses of soil microbial communities under grassland
restoration. Agriculture, ecosystems & environment, 127(3-4), pp.286-293.
71 Podrázský, V., Fulín, M., Prknová, H., Beran, F. and Třeštík, M., 2016. Changes of
agricultural land characteristics as a result of afforestation using introduced tree species.
Journal of Forest Science, 62(2), pp.72-79.
72 Podrázský, V.P.F.C., Remeš, J., Hart, V. and Moser, W.K., 2009. Production and humus
form development in forest stands established on agricultural lands–Kostelec nad
Černými lesy region. Journal of Forest Science, 55(7), pp.299-305.
73 Poeplau, C. & Don, A. Sensitivity of soil organic carbon stocks and fractions to different
land-use changes across Europe. Geoderma 192, 189–201 (2013).

145
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

74 Prévosto, B., Dambrine, E., Coquillard, P. and Robert, A., 2006. Broom (Cytisus
scoparius) colonization after grazing abandonment in the French Massif Central: impact
on vegetation composition and resource availability. Acta Oecologica, 30(2), pp.258-
268.
75 Priemé, A., Christensen, S., Dobbie, K.E. and Smith, K.A., 1997. Slow increase in rate
of methane oxidation in soils with time following land use change from arable
agriculture to woodland. Soil Biology and Biochemistry, 29(8), pp.1269-1273.
76 Rahman, M.M., Bàrcena, T.G. and Vesterdal, L., 2017. Tree species and time since
afforestation drive soil C and N mineralization on former cropland. Geoderma, 305,
pp.153-161.
77 Rodrigo-Comino, J., Martínez-Hernández, C., Iserloh, T., Cerdà, A., 2018. Contrasted
Impact of Land Abandonment on Soil Erosion in Mediterranean Agriculture Fields.
Pedosphere. https://doi.org/10.1016/S1002-0160(17)60441-7
78 Rodrigo-Comino, J., Neumann, M., Remke, A. and Ries, J.B., 2018. Assessing
environmental changes in abandoned German vineyards. Understanding key issues for
restoration management plans. Hungarian Geographical Bulletin, 67(4), pp.319-332.
79 Roldan, A., Garcia, C., Albaladejo, J., 1997. AM fungal abundance and activity in a
chronosequence of abandoned fields in a semiarid mediterranean site. Arid Soil Res.
Rehabil. https://doi.org/10.1128/IAI.01216-06
80 Romanyà, J., Cortina, J., Falloon, P., Coleman, K., Smith, P., 2000. Modelling changes
in soil organic matter after planting fast-growing Pinus radiata on Mediterranean
agricultural soils. Eur. J. Soil Sci. 51, 627–641.
81 Romero-Díaz, A., Ruiz-Sinoga, J.D., Robledano-Aymerich, F., Brevik, E.C., Cerdà, A.,
2017. Ecosystem responses to land abandonment in Western Mediterranean Mountains.
Catena. https://doi.org/10.1016/j.catena.2016.08.013
82 Rosenqvist, L., Kleja, D.B. and Johansson, M.B., 2010. Concentrations and fluxes of
dissolved organic carbon and nitrogen in a Picea abies chronosequence on former arable
land in Sweden. Forest Ecology and Management, 259(3), pp.275-285.
83 Ruecker, G., Schad, P., Alcubilla, M.M., Ferrer, C., 1998. Natural regeneration of
degraded soils and site changes on abandoned agricultural terraces in Mediterranean
Spain. L. Degrad. Dev. 9, 179–188.
84 Sciubba, L., Mazzon, M., Cavani, L., Baldi, E., Toselli, M., Ciavatta, C. and Marzadori,
C., 2021. Soil Response to Agricultural Land Abandonment: A Case Study of a Vineyard
in Northern Italy. Agronomy, 11(9), p.1841.
85 Seeber, J., Tasser, E., Rubatscher, D., Loacker, I., Lavorel, S., Robson, T.M., Balzarolo,
M., Altimir, N., Drösler, M., Vescovo, L. and Gamper, S., 2022. Effects of land use and
climate on carbon and nitrogen pool partitioning in European mountain grasslands.
Science of The Total Environment, p.153380.
86 Segura, C., Jiménez, M.N., Nieto, O., Navarro, F.B., Fernández-Ondoño, E., 2016.
Changes in soil organic carbon over 20 years after afforestation in semiarid SE Spain.
For. Ecol. Manage. https://doi.org/10.1016/j.foreco.2016.09.035
87 Smal, H. and Olszewska, M., 2008. The effect of afforestation with Scots pine (Pinus
silvestris L.) of sandy post-arable soils on their selected properties. II. Reaction, carbon,
nitrogen and phosphorus. Plant and Soil, 305(1), pp.171-187.
88 Smal, H., Ligęza, S., Pranagal, J., Urban, D. and Pietruczyk-Popławska, D., 2019.
Changes in the stocks of soil organic carbon, total nitrogen and phosphorus following
afforestation of post-arable soils: A chronosequence study. Forest Ecology and
Management, 451, p.117536.

146
CHAPTER IV: Factors driving soil carbon sequestration following agricultural land abandonment in Europe

89 Sokołowska, J., Józefowska, A., Woźnica, K. and Zaleski, T., 2020. Succession from
meadow to mature forest: Impacts on soil biological, chemical and physical properties—
Evidence from the Pieniny Mountains, Poland. Catena, 189, p.104503.
90 Spohn, M., Novák, T.J., Incze, J. and Giani, L., 2016. Dynamics of soil carbon, nitrogen,
and phosphorus in calcareous soils after land-use abandonment–A chronosequence
study. Plant and soil, 401(1-2), pp.185-196.
91 Strand, L.T., Fjellstad, W., Jackson-Blake, L. and De Wit, H.A., 2021. Afforestation of a
pasture in Norway did not result in higher soil carbon, 50 years after planting. Landscape
and Urban Planning, 207, p.104007.
92 Valverde-Asenjo, I., Diéguez-Antón, A., Martín-Sanz, J.P., Molina, J.A., Quintana, J.R.,
2020. Soil and vegetation dynamics in a chronosequence of abandoned vineyards. Agric.
Ecosyst. Environ. 301. https://doi.org/10.1016/j.agee.2020.107049
93 van der Wal, A., van Veen, J.A., Smant, W., Boschker, H.T., Bloem, J., Kardol, P., van
der Putten, W.H. and de Boer, W., 2006. Fungal biomass development in a
chronosequence of land abandonment. Soil Biology and Biochemistry, 38(1), pp.51-60.
94 Van Sundert, K., Brune, V., Bahn, M., Deutschmann, M., Hasibeder, R., Nijs, I. and
Vicca, S., 2020. Post-drought rewetting triggers substantial K release and shifts in leaf
stoichiometry in managed and abandoned mountain grasslands. Plant and Soil, 448(1),
pp.353-368.
95 Varnagirytė-Kabašinskienė, I., Žemaitis, P., Armolaitis, K., Stakėnas, V. and Urbaitis,
G., 2021. Soil Organic Carbon Stocks in Afforested Agricultural Land in Lithuanian
Hemiboreal Forest Zone. Forests, 12(11), p.1562.
96 Vesterdal, L., Rosenqvist, L., Van Der Salm, C., Hansen, K., Groenenberg, B.J. and
Johansson, M.B., 2007. Carbon sequestration in soil and biomass following afforestation:
experiences from oak and Norway spruce chronosequences in Denmark, Sweden and the
Netherlands. In Environmental effects of afforestation in north-western Europe (pp. 19-
51). Springer, Dordrecht.
97 Vopravil, J., Podrázský, V., Khel, T., Holubík, O. and Vacek, S., 2014. Effect of
afforestation of agricultural soils and tree species composition on soil physical
characteristics changes. Ekológia, 33(1), p.67.
98 Wellock, M.L., LaPerle, C.M. and Kiely, G., 2011. What is the impact of afforestation
on the carbon stocks of Irish mineral soils?. Forest Ecology and Management, 262(8),
pp.1589-1596.
99 Zeller, V., Bardgett, R.D. and Tappeiner, U., 2001. Site and management effects on soil
microbial properties of subalpine meadows: a study of land abandonment along a north–
south gradient in the European Alps. Soil Biology and Biochemistry, 33(4-5), pp.639-
649.
100 Zethof, J.H.T., Cammeraat, E.L.H., Nadal-Romero, E., 2019. The enhancing effect of
afforestation over secondary succession on soil quality under semiarid climate
conditions. Sci. Total Environ. 652, 1090–1101.
https://doi.org/10.1016/J.SCITOTENV.2018.10.235
101 Zhiyanski, M., Glushkova, M., Ferezliev, A., Menichetti, L. and Leifeld, J., 2016.
Carbon storage and soil property changes following afforestation in mountain
ecosystems of the Western Rhodopes, Bulgaria. iForest-Biogeosciences and Forestry,
9(4), p.626.
102 Zornoza, R., Mataix-Solera, J., Guerrero, C., Arcenegui, V., Mataix-Beneyto, J., 2009.
Comparison of soil physical, chemical, and biochemical properties among native forest,
maintained and abandoned almond orchards in mountainous areas of Eastern Spain. Arid
L. Res. Manag. 23, 267–282. https://doi.org/10.1080/15324980903231868

147
CHAPTER V: Conclusions and future perspectives
CHAPTER V: Conclusions and future perspectives

5.1 Conclusions
The overall aim of this PhD thesis is to generate new knowledge on the effects of ALA on soil
carbon stocks by exploring the temporal dynamics of soil organic carbon following the
cessation of agricultural activities at the field, regional, and continental scales. This thesis
provides novel insights into the capacity of European agricultural soils to recarbonize through
ecological succession, informing international ecosystem restoration policies and land
management strategies on the potential carbon benefits, costs, and challenges of post-
agricultural landscapes.

Establishing new land uses on abandoned agricultural lands is becoming increasingly attractive
as global demand for land, food, and energy intensifies. This presents important opportunities
for carbon sequestration co-benefits. A literature review was conducted to identify the foremost
proposed management strategies for abandoned agricultural lands and compare their reported
soil carbon sequestration rates, at any temporal or spatial scale (RQI). Six major categories
were identified, each with positive and negative, direct and indirect outcomes for climate
change mitigation efforts depending on site-specific factors and management objectives (Table
3). Accordingly, no single strategy is ideal in all scenarios and a combination of strategies can
address multiple rural development goals concurrently. These site-specific factors also play a
significant role in the soil carbon sequestration efficacy of each proposal in each agricultural
region of the world, resulting in high variability among rates (Figure 10). A combination of
passive and active management techniques is the most effective approach for maximizing soil
carbon sequestration over large geographic scales, while other strategies can be designed to
also promote low-carbon land use practices and fossil fuel substitution. The ecological and
rural development implications of each management strategy and new land use highlighted in
CHAPTER II informs policymakers tasked with planning the future of rural areas experiencing
ALA. To better quantify what past and present ALA implies for climate change mitigation,
temporal analyses featuring a variety of agricultural practices and crop types abandoned in
different bioclimates should be considered. However, this requires overcoming the challenge
of gathering soil data at the decadal scale and longer.

One of the global hot-pots of ALA over the last century has been the Mediterranean region of
Europe, with significant rates of ALA projected to continue. While secondary succession on
abandoned agricultural lands globally can be expected to promote SCS, the accumulation of
SOC in Mediterranean countries has been difficult to predict and is subject to multiple
competing factors. Gains, losses, and no significant changes have all been reported. Therefore,

150
CHAPTER V: Conclusions and future perspectives

field work was conducted to explore the effects of depth and time on SCS following ALA in
typical Mediterranean agricultural and secondary forest environments, and a new dataset of
chronosequences of ALA and paired-plots from peninsular Spain was generated and
synthesized to identify the potential factors responsible for the high variability in post-
agricultural SCS rates observed in the Mediterranean region (RQII). Chronosequence field
studies indicated an average SOC concentration accumulation rate of +2.3% yr–1 post-
abandonment (Figure 16); but it is a highly variable process, depending on multiple
environmental and land management factors. The highest rates of SOC accumulation post-
abandonment can be expected on lands previously used for woody crop production featuring
~13–17 ° C MAT and ~450–900 mm MAP, with the lowest rates expected on lands previously
used for annual crop production outside this climatic window. Interestingly, the secondary
forest field sites accrued 40.8 Mg C ha–1 (+172%) following abandonment but displayed greater
SOC and N depth heterogeneity than natural forests (Figure 13), demonstrating the long-lasting
impact of agricultural practices. Overall, the findings highlighted in CHAPTER III
demonstrated that ALA has produced divergent increases in SOC concentrations in peninsular
Spain. By altering the SOC accumulation rates of existing secondary forests and influencing
the locations and crop types of future ALA, precipitation and temperature changes in the
Mediterranean region will determine the SCS potential and ecological value of abandoned
agricultural lands. Regional climate change mitigation policies in Mediterranean and semi-arid
environments can consider ALA as a low-cost but long-term option best incorporated in tandem
with other multipurpose sustainable land management strategies.

At the continental scale, ALA is also a prominent land use change throughout Europe, with
several notable implications for soil health, ecosystem restoration, and transboundary rural
development planning. However, large uncertainties on the variability of post-abandonment
SCS rates (as evident in Spain, CHAPTER III) and the absolute storage potentials across
Europe hinders the development of dedicated policies leveraging the restoration benefits of
both intentional (i.e., managed restoration and direct conversions) and unintentional, unplanned
ALA. To provide critical information and new insights in support of the next generation of
European land use and climate-related policies seeking to leverage post-agricultural soils as
carbon sinks, the largest dataset ever collected on SOC stock changes specifically following
agricultural land abandonment/conversion at a continental scale was produced and synthesized.
By extracting over 800 data-pairs from published chronosequences and paired plots and
estimating 546 individual soil profiles, the potential environmental and human management

151
CHAPTER V: Conclusions and future perspectives

factors driving SCS rates following ALA in Europe were investigated (RQIII). There is a slow,
but significant, rate of SOC stock increase across Europe, at 1.28% yr–1, with an absolute rate
of 0.32 Mg ha–1 yr–1 (Figure 29). The mean relative and absolute increase amongst the data-
pairs is 32.1% and 10.5 Mg ha–1, respectively, with an average time since abandonment of 34
years.

Figure 29. Relative (a) and absolute (b) change in SOC stock (Mg ha–1) over time for the full
European dataset of ALA.
These results provide some explanation behind the regional debates on the positive, negative,
neutral SCS potential of post-agricultural soils, which have likely been confounded by other
key factors. In general, sites with low initial stock had greater potential for SOC accumulation
while sites with high initial stock are presumably closer to SOC saturation and unlikely to
exhibit large relative increases post-abandonment/conversion. Abandoned agricultural lands in
biogeographical regions featuring optimal climatic windows had higher SOC sequestration
rates, but human management factors can produce both positive and negative effects on SOC,
resulting in several strongly divergent responses to ALA (Figure 27). Past croplands had a
notably greater rate of SOC increase over time than sites that were previously used as pastures,
likely a result of lower initial SOC stocks in croplands compared to pastures. Sites that
underwent natural ecological succession exhibited a greater rate of change in SOC stock
relatively compared to sites that were actively restored or converted to new vegetation land
covers, for example through tree planting practices. CHAPTER IV suggests that abandoned
croplands with low initial SOC stock and managed through natural succession would show the
greatest SOC accrual in Europe, while fertile pastures that are actively converted (e.g.,
afforested) would result in the lowest increases in SOC, or even losses. The variability in post-

152
CHAPTER V: Conclusions and future perspectives

abandonment/conversion SOC dynamics must be considered in sustainable land use planning


that strives to incorporate the positive ecological and climate change mitigation implications
of ALA, taking into account site-specific conditions and past and present land management
histories to avoid detrimental impacts for soil health and lost opportunities for climate change
mitigation.

Overall, the findings presented in this PhD thesis helps inform ecosystem restoration policies
and land management strategies on the potential soil carbon benefits, costs, and challenges of
post-agricultural landscapes. While the reported SCS rates on abandoned/converted
agricultural lands are generally positive for all proposed land management strategies
(CHAPTER II), this thesis demonstrated an overarching trend defining the temporal responses
of SOC stocks to ALA during revegetation over large geographic scales: divergencies
depending on initial carbon stock, past land use, past crop type, restoration management
regime, and climate/biogeographical variables (CHAPTER III & IV). The high variability in
post-abandonment/conversion SOC temporal dynamics must be considered in sustainable land
use planning that strives to incorporate the positive ecological and climate change mitigation
implications of ongoing ALA at regional and continental scales, taking into account site-
specific conditions and past and present land management factors.

153
CHAPTER V: Conclusions and future perspectives

5.2 Future perspectives


Maximizing natural climate solutions is becoming increasingly critical as atmospheric CO2
concentrations continue to rise (Friedlingstein et al., 2022; Griscom et al., 2017). As another
an indication of the level of urgency, Matthews et al., (2022) argues it is no longer even a
question of carbon sequestration permanence: even temporary and later reemitted nature-based
carbon removal should be pursued to keep global temperature rise below 2° C. Aside from
emissions reductions, sustainable and climate-smart land management is one of our most
important objectives (IPCC, 2019). Mitigating the negative impacts of climate change by
returning carbon to depleted soils, in particular, will require exploring all available avenues
(Bossio et al., 2020).

The combined results from this thesis present a promising outlook for leveraging ongoing ALA
processes for soil carbon sequestration, and also ecosystem restoration more broadly (Yang et
al., 2020). But despite calls for protection (Poore, 2016), unprotected abandoned agricultural
lands often lead to recultivation and lost opportunities for climate change mitigation (Crawford
et al., 2022). I argue we must further explore integrating strategic post-agricultural landscapes
into land use planning, while protecting existing abandoned agriculture lands whenever
feasible (i.e., avoiding conflicts with food production and land rights by focusing especially on
uncontested abandoned lands (Xie et al., 2020)). This can help diversify the global land carbon
sink and support the objectives of the UN Decade of Ecosystem Restoration (2021-2030)
(Abhilash, 2021; Aronson et al., 2020). To that end, I propose three next-steps:

1. Expanding the dataset to all continents. The inferential potential of assembled


chronosequences and paired-plots has not yet been realized. Just like in Europe, there
exists thousands of un-synthesized, disparate data-pairs of post-agricultural sites in
other continents, with hundreds more published every year. A large-scale global
synthesis is needed along the lines of Cook-Patton et al., (2020) or Veldkamp et al.,
(2020), both of which featured large collections of time stamped data-pairs on soil
properties following land use change from agriculture. Special attention must be paid
to underrepresented areas in global change science, namely Africa, South America, and
most of Asia. The importance of this task is twofold: 1) chronosquences/paired-plots
are logistically superior to long-term experimental plots with repeated measurements
and should therefore be exploited as much as possible to advance urgent global change
science; and, 2) archiving this kind of data for future research lowers the risk of data
loss, repurposes valuable past investments, and increases data source diversity which

154
CHAPTER V: Conclusions and future perspectives

improves model accuracy (i.e., avoiding data “siloing” by repeatedly analyzing the
most popular or available datasets).
2. Improving post-agricultural soil carbon sequestration modelling. The high
variability in SOC temporal responses found across Europe are a function of the
complexity of soils and their interactions with the environment and human activities.
Nevertheless, dedicated efforts need to be made to better predict how SOC stocks of a
given plot of land will respond to ALA, expanding from linear regression analyses to
more appropriate estimations of SOC saturation functions. Just as it is important to
produce more accurate soil carbon sequestration rates for different categories of
abandoned agricultural lands, it is also necessary to predict their specific equilibrium
parameters (i.e., time to equilibrium, equilibrium SOC stock value). In this exercise, it
will also be important to account for the different SOC pools (e.g., MAOM and POM),
which will have different rates of change than SOM as a whole. The rates of change for
these pools are normally assumed in equilibrium models because they are not
represented well in most datasets. By extracting, synthesizing, and repurposing time-
stamped SOC data from any published study of ALA, no matter the original research
angle, we can more robustly benchmark and validate models of successional carbon
dynamics.
3. Combining more robust, data-informed models with next-generation ALA spatial
estimates. After improving temporal SOC data quality and quantity, the lack of
accurate, high-resolution ALA spatial estimates at the global scale should be addressed.
Current state-of-the-art maps only cover the past few decades, and struggle to capture
abandoned pastures and smaller-sized field plots. Fortunately, new sensors are now
available with higher spatial, temporal, and spectral resolution allowing for improved
monitoring of active agricultural lands, which by extension improves our abilities to
identify and map post-agricultural landscapes such as abandoned lands. For example,
Sentinel-1 and Sentinel-2, together with Landsat, provide a denser time series at the
global scale than what has been previously available. There is also new, very high-
resolution (e.g., < 5m) commercial satellite imagery now available that can be used to
isolate and process smaller agricultural plots (e.g., PlanetScope). The large volumes of
data generated by these new sensors can also now be processed more rapidly and with
more accessibility than ever before using free cloud computation resources like Google
Earth Engine. Ongoing agricultural classifications via ground truthing at the global
scale will help validate these newly produced, high-quality agricultural land cover

155
CHAPTER V: Conclusions and future perspectives

maps. And with the continual improvement of machine learning approaches at our
disposal, we can anticipate the first reliable and detailed global maps of past and present
abandoned agricultural lands, and their durations of existence, in the coming years.

I encourage a concerted effort by researchers, policy-makers, and land managers to constrain


the impact of ALA on the global land carbon sink as soon as possible. This PhD thesis has
served as a springboard for the first global data collection initiative focused on the temporal
effects of ALA on soil carbon stocks. Our team of data collectors, led by myself for the past
1.5 years, have expanded the European dataset to every inhabited continent of the world (Figure
30). Our goal is to provide an antithesis publication to the excellent work by Sanderman et al.,
(2017) by quantifying how ALA has helped us pay back agriculture’s longstanding soil carbon
debt, putting carbon back where it belongs.2

Figure 30. Global distribution of chronosequences and paired-plot data-pairs (n = 3460)


collected by myself and our team of MSc and PhD student colleagues in the inter-university
project created from this PhD thesis.

2
www.ted.com/talks/stephen_bell_retiring_farmlands_to_put_carbon_back_where_it_belongs

156
CHAPTER V: Conclusions and future perspectives

5.3 References
1. Abhilash, P. C. (2021). Restoring the Unrestored: Strategies for Restoring Global Land during the UN Decade
on Ecosystem Restoration (UN-DER). Land, 10(2), 201. https://doi.org/10.3390/land10020201

2. Aronson, J., Goodwin, N., Orlando, L., Eisenberg, C., & Cross, A. T. (2020). A world of possibilities: six
restoration strategies to support the United Nation’s Decade on Ecosystem Restoration. Restoration Ecology,
28(4), 730–736. https://doi.org/10.1111/rec.13170

3. Bossio, D. A., Cook-Patton, S. C., Ellis, P. W., Fargione, J., Sanderman, J., Smith, P., Wood, S., Zomer, R.
J., von Unger, M., Emmer, I. M., & Griscom, B. W. (2020). The role of soil carbon in natural climate
solutions. Nature Sustainability. https://doi.org/10.1038/s41893-020-0491-z

4. Cook-Patton, S. C., Leavitt, S. M., Gibbs, D., Harris, N. L., Lister, K., Anderson-Teixeira, K. J., Briggs, R.
D., Chazdon, R. L., Crowther, T. W., Ellis, P. W., Griscom, H. P., Herrmann, V., Holl, K. D., Houghton, R.
A., Larrosa, C., Lomax, G., Lucas, R., Madsen, P., Malhi, Y., … Griscom, B. W. (2020). Mapping carbon
accumulation potential from global natural forest regrowth. Nature, 585(7826), 545–550.
https://doi.org/10.1038/s41586-020-2686-x

5. Crawford, C. L., Yin, H., Radeloff, V. C., & Wilcove, D. S. (2022). Rural land abandonment is too ephemeral
to provide major benefits for biodiversity and climate. Science Advances, 8(21).
https://doi.org/10.1126/sciadv.abm8999

6. Friedlingstein, P., Jones, M. W., O’Sullivan, M., Andrew, R. M., Bakker, D. C. E., Hauck, J., le Quéré, C.,
Peters, G. P., Peters, W., Pongratz, J., Sitch, S., Canadell, J. G., Ciais, P., Jackson, R. B., Alin, S. R., Anthoni,
P., Bates, N. R., Becker, M., Bellouin, N., … Zeng, J. (2022). Global Carbon Budget 2021. Earth System
Science Data, 14(4), 1917–2005. https://doi.org/10.5194/essd-14-1917-2022

7. Griscom, B. W., Adams, J., Ellis, P. W., Houghton, R. A., Lomax, G., Miteva, D. A., Schlesinger, W. H.,
Shoch, D., Siikamäki, J. v., Smith, P., Woodbury, P., Zganjar, C., Blackman, A., Campari, J., Conant, R. T.,
Delgado, C., Elias, P., Gopalakrishna, T., Hamsik, M. R., … Fargione, J. (2017). Natural climate solutions.
Proceedings of the National Academy of Sciences, 114(44), 11645–11650.
https://doi.org/10.1073/pnas.1710465114

8. IPCC. (2019). IPCC Special Report on Climate Change, Desertification, Land Degradation, Sustainable Land
Management, Food Security, and Greenhouse gas fluxes in Terrestrial Ecosystems. Summary for
Policymakers.

9. Matthews, H. D., Zickfeld, K., Dickau, M., MacIsaac, A. J., Mathesius, S., Nzotungicimpaye, C.-M., & Luers,
A. (2022). Temporary nature-based carbon removal can lower peak warming in a well-below 2 °C scenario.
Communications Earth & Environment, 3(1), 65. https://doi.org/10.1038/s43247-022-00391-z

10. Poore, J. A. C. (2016). Call for conservation: Abandoned pasture. Science, 351(6269), 132.

11. Sanderman, J., Hengl, T., & Fiske, G. J. (2017). Soil carbon debt of 12,000 years of human land use.
Proceedings of the National Academy of Sciences. https://doi.org/10.1073/pnas.1706103114

12. Veldkamp, E., Schmidt, M., Powers, J. S., & Corre, M. D. (2020). Deforestation and reforestation impacts on
soils in the tropics. Nature Reviews Earth & Environment, 1(11), 590–605. https://doi.org/10.1038/s43017-
020-0091-5

13. Xie, Z., Game, E. T., Hobbs, R. J., Pannell, D. J., Phinn, S. R., & McDonald-Madden, E. (2020). Conservation
opportunities on uncontested lands. Nature Sustainability, 3(1), 9–15. https://doi.org/10.1038/s41893-019-
0433-9

14. Yang, Y., Hobbie, S. E., Hernandez, R. R., Fargione, J., Grodsky, S. M., Tilman, D., Zhu, Y.-G., Luo, Y.,
Smith, T. M., Jungers, J. M., Yang, M., & Chen, W.-Q. (2020). Restoring Abandoned Farmland to Mitigate
Climate Change on a Full Earth. One Earth, 3(2), 176–186. https://doi.org/10.1016/j.oneear.2020

157

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy