You are on page 1of 77

DDeeppaar trm

tmeennt toof fBBuuilitltEEnnvvirio


ronnmmeennt t
Sa
S

RReem mootteellyy sseennsseedd


a rra
aAA ll ii b

mmoonniittoorriinngg ooff llaanndd ssuurrffaaccee


baak
khh ss h

aallbbeeddoo aanndd eeccoossyysstteem m


R
Reem h ii

ddyynnaam miiccss
mootte
ellyy s
seen
nsse
ed

SSaara
raAAlilbibaakkhhsshhi i
dm
mo
Aalto University
Aalto University on
niitto
orriin
nggo
off lla
annd
dssu
urr ffa
acce
eaallb
beed
do
oaan
ndde
ecco
ossyys
stte
emmd
dyyn
naam
miic
css

DD
OO CC TT
OORRAALL
DD
I SI S
SSEERRTT
AATT
IOIO
NNSS

22
0022
00
Aalto University publication series
DOCTORAL DISSERTATIONS 183/2020

Remotely sensed monitoring of land


surface albedo and ecosystem
dynamics

Sara Alibakhshi

A doctoral dissertation completed for the degree of Doctor of


Science (Technology) to be defended, with the permission of the
Aa l t o U n i v e r s i t y S ch o ol o f E n g i n ee r i n g, r e m o t e c on n e c t i on v i a Zo om
on 04 December 2020 at 12pm.

Aalto University
School of Engineering
Department of Built Environment
Geoinformatics
Supervising professor
Professor Miina Rautiainen, Aalto University, Finland

Preliminary examiners
Professor Jan Verbesselt, Wageningen University, the Netherlands
Professor Aku Riihelä, Finnish Meteorological Institute, Finland

Opponent
Professor Mikko Vastaranta, University of Eastern Finland, Finland

Aalto University publication series


DOCTORAL DISSERTATIONS 183/2020

© 2020 Sara Alibakhshi

ISBN 978-952-64-0129-4 (printed)


ISBN 978-952-64-0130-0 (pdf)
ISSN 1799-4934 (printed)
ISSN 1799-4942 (pdf)
http://urn.fi/URN:ISBN:978-952-64-0130-0

Unigrafia Oy
Helsinki 2020 N
SWA ECO
C
DI

LA
NOR

BEL

Finland
Printed matter
Printed matter
1234 5678
4041-0619
Abstract
Aalto University, P.O. Box 11000, FI-00076 Aalto www.aalto.fi

Author
Sara Alibakhshi
Name of the doctoral dissertation
Re m o t e l y s e n s e d m o n i t o r i n g o f l a n d s u r f a c e a l b e d o a n d e c o s y s t e m d y n a m i c s
P u b l i s h e r School of Engineering
U n i t Department of Built Environment
S e r i e s Aalto University publication series DOCTORAL DISSERTATIONS 183/ 2020
F i e l d o f r e s e a r c h Geoinformatics
M a n u s c r i p t s u b m i t t e d 6 June 2020 D a t e o f t h e d e f e n c e 4 Decembe r 2020
P e r m i s s i o n f o r p u b l i c d e f e n c e g r a n t e d ( d a t e ) 27 October 2020 L a n g u a g e English
Monograph Article dissertation Essay dissertation
Abstract
The earth is under unprecedented pressure, which is refl ected in rapid ecosystem changes
around the globe. Over just the past three decades, the earth has lost over 178 million
hectares of its forests. The rapidly growing evidence of the loss of resilience in ecosystems
(i.e., recovery from disturbances slows down) due to climate change has become a global
concern. Our limited knowledge of ecosystem dynamics and their key parameters, such as
albedo, has also hindered our ability to manage ecosystems appropriately.
The main aim of t his d is s e rt at ion is t o cont ribu t e t o e l u cid at ing e c os y s t e m d y namic s by e xp l oit ing
remotely sensed satellite data. Additionally, this dissertation aims to explore the dynamics of
albedo (refl ectivity) in response to forest structure (forest density, tree cover, and leaf area index)
variations and forest disturbances (fi re and drought). To address specifi c research questions
discussed throughout this dissertation, various study sites extending from local to global scales are
considered.
Th e re s u l t s s h o w e d t h at u s i n g a n a p p ro p ri at e e c o s y s t e m s t at e va ri a b l e t h at c a n re p re s e n t t h e s t at e
of an ecosystem makes it is possible to achieve a reliable and timely evaluation of wetlands or
forests state change. To that end, a new remotely sensed index called the modified vegetation water
ratio (MVWR) was developed which improved the ability to understand the dynamics of wetlands.
A new approach was also developed based on incorporating local spatial autocorrelation. The
effi ciency of this approach in measuring the state of drought-affected forests was demonstrated.
To provide a convenient implementation of the presented approaches, a new R package, "stew",
was developed.
Investigating the dynamics of forest albedo on disturbances revealed that precipitation and burn
severity only weakly explained the temporal dynamics of albedo. In contrast, it was shown that the
number of fi re events and the leaf area index strongly explained temporal albedo dynamics.
According to the results, although fi res could lead to abrupt decreases in the temporal dynamics
of albedo, droughts caused abrupt increases in the temporal dynamics of albedo.
Finally, a global-scale study was conducted to explore the links between forest structure and albedo
during the peak growing season. The results demonstrated that forest structure might significantly
explain albedo in most of the forests around the world. It was also found that the response of the
s h o rt w ave a l b e d o ( 300– 5 000 nm) t o t h e va ri at i o ns o f t h e l e a f are a i nd e x w a s a l w ay s p o s i t i ve . Th e
first map representing the links between forest structure and albedo was provided which highlights
the importance of the role of forests in modulating albedo on a global scale. It is expected that the
implications of the results of this dissertation contribute to future climate mitigation plans, the
monitoring of ecosystem dynamics, and sustainable development in general.
K e y w o r d s ecosystem dynamics, land surface albedo, disturbances, forest mortality, critical
transition, satellite images
I S B N ( p r i n t e d ) 978-952-64-0129-4 I S B N ( p d f ) 978-952-64-0130-0
I S S N ( p r i n t e d ) 1799-4934 I S S N ( p d f ) 1799-4942
L o c a t i o n o f p u b l i s h e r Helsinki L o c a t i o n o f p r i n t i n g Helsinki Y e a r 2020
P a g e s 212 u r n http://urn.fi /URN:ISBN: 978-952-64-0130-0
Acknowledgments

Acknowledgments

The time for conducting this PhD research at Aalto University has been inspir-
ing and enjoyable. This dissertation has not been finished without the support
of my supervision team, colleagues, friends and family members to whom I am
deeply grateful.

First and foremost, I would like to express my sincerest appreciation to my


supervisor, Prof. Miina Rautiainen who always supported me patiently and
guided me to get this dissertation done. Before starting my PhD studies, she
supported me to write a proposal for CIMO program that was funded for 9
months, thus, I moved to Finland, and started my research in her group in the
Geoinformatics department at the Aalto University. Then, she supported me to
write a proposal and apply for a 4-year PhD fellowship at Aalto University that
was successful that staggeringly influenced my research career. I am greatly in-
debted to you, Miina, I'll never forget the first day I arrived in Finland and you
picked me up from the airport and made my settlement easier. Thanks for giving
me this opportunity to join your group and all the constructive criticism, and
the incredible patience to guide me through my research. I was even more grate-
ful when I read a sentence from the article you had shared with me about fin-
ishing a PhD while everything in life is in balance, that was “someone cares
enough to tell you the truth instead of taking the easier road of smiling and nod-
ding”.

I would like to thank Dr Aarne Hovi who collaborated in significant parts of


my studies. We had innumerable meetings and discussions about the details,
debates, and analyses. I am deeply gloried to work with you and learn many
things from you. Your frequent constructive comments and endless remarks al-
ways resulted in fruitful discussions with great outcomes where you also gener-
ously shared your knowledge about the technical aspects of my research work.
You always inspired me on how to be a good researcher. You have been more
than a colleague to me, and it is my pleasure to know you at this stage of my life.

I really appreciate the amazing supports of Prof. Thomas Crowther (Tom).


Tom, I had the opportunity to visit your lab during my three-month stay in Swit-
zerland. Through my visit, I enjoyed being at ETH, and the city of Zurich, learn-
ing from excellent scientific interactions in your lab. You helped me to very
smoothly integrate into your research group and I received the same support as
your doctoral students.

I extend true thanks to my other collaborators in parts of my studies, Prof.


Thomas Geron from the University of Twente in the Netherlands, and Dr Babak
Naimi from the University of Helsinki. Thomas, it was really constructive and
1
enjoyable to meet you in person when you visited Helsinki to discuss the details
of my first publication. Your insights on how to shape the article were really
helpful. I also appreciate the pre-examiners of this dissertation, Prof. Jan
Verbesselt and Prof. Aku Riihelä for their evaluation of this dissertation.

During my PhD studies, I had great discussions with my colleagues. Dr Titta


Majasalmi, I benefited from your concrete comments and questions to develop
my fourth paper. Juola Jussi, Dr Wang Di, Schraik Daniel, and Forsström Petri,
I enjoyed tremendously talking with you in coffee breaks and our friendly work-
ing office. I would like to also thank, Dr Hadi Hadi, who was the former PhD
candidate and a successful and diligent member of our team. Hadi, your fruitful
insights and discussion, especially in the second paper of my dissertation, lent
a hand in improving it. I acknowledge Dr Terhikki Manninen for her comments
on the second article of this dissertation. I extend true thanks to Dr Bakhtiar
Fattahi (Assistant Professor, Faculty of Natural Resources and Environment,
Malayer University, Iran) for sharing the reports on forest fires, and also Dr
Shahram Ahmadi (Senior expert at Department of natural resources in Fars,
Iran), Ahmad Hossein Zade (MSc, Faculty of Forestry, Gheir Entefai University,
Ilam, Iran) and Mohammad Jabari (MSc, Faculty of Environmental Science,
IAU, Tehran, Iran) for background information on the second paper of this dis-
sertation.

I would like to also express my appreciation and love to my wonderful par-


ents, my brother, Amin, and my sister, Saedeh. Thanks for your invaluable sup-
port, encouragement, and endless love.

Helsinki, 29 October 2020

Sara Alibakhshi

2
Contents

Contents

Acknowledgments ....................................................................................... 1
Contents ....................................................................................................... 3
List of abbreviations and symbols .............................................................. 5
List of publications ...................................................................................... 7
Author’s contribution .................................................................................. 8
1. Introduction ..................................................................................... 9
General framework ...................................................................... 9
The necessity for understanding ecosystem dynamics ............ 10
The necessity of understanding land surface albedo ................ 11
The application of remote sensing in the monitoring of ecosystems
12
Objectives of the dissertation .................................................... 13
An overview of the publications ................................................ 14
2. Literature review ............................................................................ 16
Identifying ecosystem dynamics ............................................... 16
2.1.1. Dynamic system theory.......................................................... 16
2.1.2. Ecosystem state variables .......................................................17
2.1.3. Methods to identify ecosystem dynamics ............................. 18
Albedo dynamics ........................................................................ 19
2.2.1. Mechanisms underlying albedo variations in forests ..........20
2.2.2. Forest albedo variations upon disturbances ........................ 21
2.2.3. Statistics to measure the dynamics of albedo....................... 22
3. Materials ......................................................................................... 24
Description of study sites .......................................................... 24
Data acquisition ......................................................................... 26
Data extraction and quality assessment ................................... 28
4. Methods .......................................................................................... 29
Overview of the applied methods and tools ............................. 29
Monitoring ecosystem dynamics .............................................. 31
4.2.1. Temporal early warning signals (TEW) ................................ 31

3
4.2.2. SPT analysis in forests ........................................................... 33
4.2.3. Spatio-temporal analyses of early warning signals ............. 34
Albedo dynamics ........................................................................ 35
4.3.1. Black sky albedo ..................................................................... 35
4.3.2. The temporal dynamics of albedo ......................................... 36
4.3.3. Albedo and forest structure ................................................... 37
5. Results ............................................................................................ 40
Monitoring ecosystem dynamics .............................................. 40
5.1.1. Temporal analysis of early warning signals (TEW) ............. 40
5.1.2. SPT, STG, and STM analysis ................................................. 42
The temporal dynamics of forest albedo .................................. 44
Albedo and forest structure....................................................... 47
6. Discussion ...................................................................................... 50
Monitoring ecosystem dynamics .............................................. 50
6.1.1. Ecosystem state variables ...................................................... 50
6.1.2. Methods of identifying ecosystem dynamics ....................... 50
The temporal dynamics of albedo.............................................. 51
Albedo modulation by forest structure .................................... 52
Uncertainties and recommendations for future studies ......... 54
The fusion of knowledge about ecosystem state dynamics theory with
albedo dynamics .................................................................................... 56
7. Summary .........................................................................................57
8. References ...................................................................................... 59
9. Publications.................................................................................... 72

4
List of abbreviations and symbols

List of abbreviations and symbols

ACF autocorrelation function

BFAST break detection for additive season and trend

br boreal

BRDF bidirectional reflectance distribution function

DB deciduous broadleaf

DN deciduous needleleaf

dNBR difference normalized burn ratio

EB evergreen broadleaf

EN evergreen needleleaf

ENSO El Niño/Southern Oscillation

FIRMS fire information for resource management system

GAM generalized additive model

GEE Google Earth engine

IPCC intergovernmental panel on climate change

JRC Joint Research Center

LAI leaf area index

NBR normalized burn ratio

NDVI normalized difference vegetation index

NDWI normalized difference water index

me mediterranean

Mixed mixed forests

MNDWI modified normalized difference water index


5
MVWR modified vegetation water ratio

MODIS moderate resolution imaging spectroradiometer

SD standard deviation

SK skewness

SRTM shuttle radar topography mission

STG spatio-temporal analysis of Geary’s c

STM spatio-temporal analysis of Moran’s I

SWIR shortwave infrared

SZA solar zenith angle

tem temperate

TEW temporal analysis of early warning signal

tr tropical

VCF vegetation continuous fields

WS woody savannah forests

6
List of publications

This doctoral dissertation consists of a summary and of the following publica-


tions which are referred to in the text by their numerals:

Publication I: Alibakhshi, S., Groen, T.A., Rautiainen, M., and Naimi, B.,
2017. Remotely-sensed early warning signals of a critical transition in a wetland
ecosystem. Remote sensing, 9(4), p.352. https://doi.org/10.3390/rs9040352

Publication II: Alibakhshi, S. Spatio-temporal analysis of remote sensing


images provides early warning signals of forest mortality. Submitted to Science
of the Total Environment on 22nd of May 2020 and revised according to pre-
examination comments

Publication III: Alibakhshi, S., Hovi, A., and Rautiainen, M., 2019. Tem-
poral dynamics of albedo and climate in the sparse forests of Zagros. Science of
the Total Environment, 663, pp.596-609. https://doi.org/10.1016/j.sci-
totenv.2019.01.253

Publication IV: Alibakhshi, S., Naimi, B., Hovi, A., Crowther, T.W., and
Rautiainen, M., 2020. Quantitative analysis of the links between forest structure
and land surface albedo on a global scale. Remote Sensing of Environment, 246,
p.111854. https://doi.org/10.1016/j.rse.2020.111854

7
Author’s contribution

Publication I: Remotely sensed early warning signals of a critical transition in


a wetland ecosystem.

Sara Alibakhshi was responsible for the main analysis and had a leading role in
writing the paper and analyzing the results. She and Babak Naimi conceived and
designed the project. All the co-authors participated in commenting and writ-
ing.

Publication II: Spatio-temporal analysis of remote sensing images provides


early warning signals of forest mortality.

Sara Alibakhshi was the sole-author in this work.

Publication III: Temporal dynamics of albedo and climate in the sparse for-
ests of Zagros.

Sara Alibakhshi carried out the analysis and had a leading role in writing the
paper. She and Miina Rautiainen designed the study. All authors participated in
writing the manuscript.

Publication IV: Quantitative analysis of the links between forest structure


and land surface albedo on a global scale.

Sara Alibakhshi had a leading role in planning the research objectives, devel-
oping the methods, analyzing the data, and preparing the manuscript. Thomas
W Crowther contributed to discussing the research idea. Miina Rautiainen,
Aarne Hovi, and Babak Naimi contributed to preparing the manuscript.

8
Introduction

1. Introduction

General framework

Various disturbances increasingly affect the resilience of ecosystems under the


effect of rapid global changes (Adams et al. 2010; Vicente-Serrano et al. 2020).
The resilience of ecosystems is a crucially important property in evaluating the
stability of ecosystems and is defined as the ability of an ecosystem to recover
upon a disturbance (Scheffer 2009). Measuring the resilience of an ecosystem
allows us to characterize the dynamics of that ecosystem and evaluate the degree
to which the ecosystem is approaching a critical transition or abrupt change
based on the “dynamic system theory” (subchapter 1.2). The application of the
dynamic system theory in identifying the state of ecosystems and its current
limitations are discussed in subchapter 2.1.
These disturbances are also continuously altering the key biogeophysical var-
iables of ecosystems, such as land surface albedo. Albedo is defined as the ratio
of the radiant flux (amount of energy through a surface) for light reflected by a
unit surface area into the view hemisphere to the illumination radiant flux when
the surface is illuminated with a parallel beam of light from a single direction
(Schaepman-Strub et al. 2006). The dynamics of albedo in a forest are defined
as the forces or properties (either natural parameters, such as forest structure
variations, or catastrophic events, such as droughts and fires), which drive
changes in forest albedo. Monitoring albedo dynamics provides valuable infor-
mation about the amount of radiant energy absorption and reflection by the sur-
face of the earth (Betts 2000; Bright et al. 2017). The importance of albedo dy-
namics is elaborated in subchapter 1.3., and the knowledge gaps are discussed
in subchapter 2.2.
In order to measure and monitor remotely sensed land surface albedo and
ecosystem dynamics, the present dissertation takes advantage of optical satellite
images in two different ecosystems, including forests and wetlands, for which a
number of specific research challenges are addressed. It allows us to improve
the remote sensing of ecosystem dynamics and also understand land surface al-
bedo dynamics as one of the main focal points of the carried-out research during
my doctoral studies (subchapter 1.5 and subchapter 1.6, Fig. 1).
Notably, although the primary focus of the present dissertation was on wet-
land and forest ecosystems, the broader term of “ecosystem” is selected in the

9
title of this dissertation because the knowledge, methods, and approaches pre-
sented in the current dissertation can be applied to other ecosystems as well.

Figure 1. A conceptual framework, representing the components of the present study, which
highlights how different scientific objectives are linked. This dissertation explored how disturb-
ances (i.e., fires and droughts) affect the dynamics of an ecosystem measured by key variables,
such as albedo. The main research questions of the different publications are presented in the
oval-shaped boxes, and the main objectives of the publications are presented in the square-
shaped boxes.

The necessity for understanding ecosystem dynamics

The detrimental effects of climate-induced disturbances on forests have reached


an alarming global level (Allen et al. 2010; van der Werf et al. 2017; Vicente-
Serrano et al. 2020). The earth has lost 178 million hectares of forests between
1990 and 2020 due to natural and anthropogenic changes (FAO 2020). The
areas burned by wildfires have increased globally by a remarkable ~37%
between 1997 to 2016 (van der Werf et al. 2017).
A direct consequence of the increasing amount of such disturbances is the
growing global evidence of critical transition occurrences in ecosystems
(Ratajczak et al. 2018; McDowell et al. 2020), implying a higher potential to
threaten our planet. Over the past few decades, the earth has experienced sev-
eral massive extinctions of various species in different regions (Barnosky et al.
2011). Such critical transitions are likely to happen in the near future as well
(Scheffer 2009). Therefore, research priorities to advance our understanding of
ecosystem dynamics upon climate-induced disturbances have been suggested
(Turner et al. 2020).
Monitoring ecosystem dynamics can enable us to measure the resilience loss
that is of significant importance in a timely manner in order to generate early

10
Introduction

warning signals of a forthcoming critical change. Furthermore, monitoring the


state of an ecosystem purveys essential insights into the functioning of ecosys-
tems. This enables us to approach the sustainable development milestones (Lu
et al. 2015) which, as a modern life challenge, have been the main focal point of
a number of recent international agreements and panels, e.g., the intergovern-
mental panel on climate change (IPCC) in 2018 and reducing deforestation and
forest degradation (REDD+) plans (Angelsen 2009). A review of recent studies,
as well as further discussions on knowledge gaps in monitoring of ecosystem
dynamics, are elaborated in subchapter 2.1.

The necessity of understanding land surface albedo

Various national and international campaigns, agreements, and panels, such as


the IPCC, have been conducted to protect our planet from climate-induced dis-
turbances. For example, the IPCC suggested restoring one billion hectares of
forests for the sake of limiting global warming to a 1.5°C increase by 2050 (IPCC
2018). The harvesting of forests is one of the potential resources of climate
warming in some regions through critical biogeochemical variables such as land
surface albedo (Alkama & Cescatti 2016; Naudts et al. 2016). Those forests that
have the potential to decrease albedo along with other factors can offset the
cooling effects of forests in some regions (Sagan et al. 1979; Betts 2000). A study
demonstrated the failure of climate mitigation plans due to ignoring biogeo-
physical parameters, such as albedo, in forest management practices (Naudts et
al. 2016).
Albedo has a crucial role in determining the amount of energy absorption by
an ecosystem that is controlled by ecosystem components, atmosphere compo-
sition (e.g., aerosols, water vapor), and cloudiness (Betts 2000; Chambers &
Chapin 2002; Liang 2005). According to the report of the global climate observ-
ing system (GCOS), albedo is an essential variable in climate studies (GCOS
2003, 2006). The amount of absorbed energy controls various important pro-
cesses, such as the rate of photosynthesis in vegetation (Knyazikhin et al. 1998).
It also affects land surface temperature and the rate of variations in evapotran-
spiration (Su 2002; Bright et al. 2014, 2017).
It is assumed that under the same solar radiation and snow-free condition,
land surface albedo can be considered relatively stable unless a disturbance oc-
curs (Liang 2005). Monitoring of spatio-temporal albedo variations can be used
to evaluate the total effects of disturbances, such as fires, droughts, fungus, and
beetle attacks, in energy balance systems (e.g., Ghulam et al. 2007). Determin-
ing the impacts disturbances have on albedo is especially important when con-
sidering the increasing rate of forest mortality can in itself amplify albedo
changes (O’Halloran et al. 2012). This highlights the urgent need to gain a sound
understanding of determining the disturbance impacts on albedo (5th IPCC
report, Stocker et al. 2013). Nevertheless, our understanding of albedo dynam-
ics and its controlling mechanisms in terrestrial ecosystems is still quite shal-
low. It indeed suggests that an appropriate consideration of albedo dynamics,

11
including the effects of disturbances on albedo and the controlling mechanisms,
is needed.

The application of remote sensing in the monitoring of


ecosystems

Remote sensing provides cost-efficient datasets from several sensors, wave-


lengths, and spatio-temporal resolutions. The broad diversity of available sen-
sors and products with different spatio-temporal resolutions enables us to per-
form a wide range of analysis, mapping, and monitoring of various processes in
ecosystems. Remote sensing data also allows us to monitor the state of an eco-
system and its processes on a large scale. Considering that monitoring an eco-
system typically requires enriched multidimensional data in terms of spatial,
temporal, and spectral resolution, remote sensing data is of remarkably signifi-
cant importance, especially as gathering multidimensional data through field-
work can be quite impractical in harsh ecosystems, e.g., rugged terrain or desert,
or in hard to reach locations. For ecosystems in the Middle East, for example,
the long-term ground truth data are rarely available. For all such cases, our un-
derstanding of ecosystem dynamics and climate change strongly relies on satel-
lite data.
Among all the diverse remote sensing data, this study used only freely availa-
ble optical satellite images, including Landsat and moderate resolution imaging
spectroradiometer (MODIS) images, which are ideal for monitoring purposes.
Optical sensors can be applied to capture the earth’s surface reflectance across
different broadbands of visible, near-infrared, and shortwave radiation. Reflec-
tance is defined as a directional quantity that describes the relative brightness
of a spatial unit surface (Riihelä 2013). These bands cover the spectral charac-
teristics of different processes, as well as the characteristics of a spatial unit.
Since 1972, the different sensors of Landsat, including Multispectral Scanner
(MSS), Thematic Mapper (TM), Enhanced Thematic Mapper Plus (ETM+), and
Operational Land Imager (OLI), have been available to be exploited in order to
measure land surface information in multispectral bands at a 30-m spatial res-
olution. Landsat archives provide the most prolonged time series of multispec-
tral data with a high spatial resolution and global coverage. A significant ad-
vantage of Landsat-type data is the visibility of many ecosystem changes that
can only be tracked using fine spatial resolution. While Landsat is one of the
highest spatial resolution long-term sensors available, the revisit times to mon-
itor a specific site on the Earth can be very long. Although MODIS images have
a coarser spatial resolution, they offer a far more constant monitoring capability
in the temporal sense. More specifically, since 2000, the MODIS sensors, Aqua
and Terra, measure land surface information at various spatial resolutions, such
as 250-, 500-, and 1000-m. Furthermore, MODIS purveys the freely available
data with the highest temporal resolution at a medium spatial resolution. Mon-
itoring temporal dynamics of ecosystems, as well as land surface albedo, in the

12
Introduction

context of this dissertation, is very much dependent on high-frequency data,


which makes MODIS data an ideal tool for that purpose.
One of the most essential remotely sensed variables, used throughout in this
dissertation, is albedo, which was introduced in the previous chapter. The new
generation of techniques to achieve remotely sensed land surface albedo data
with an appropriate spatial resolution has emerged with the idea of attaching
sensors to aircraft (e.g., Kubelka & Munk 1931). Thereafter, a wealth of satellite
remote sensing products has begun to provide albedo data, especially during the
last two decades, such as the global land surface satellite (GLASS) (Liang et al.
2013), GlobalAlbedo (Lewis et al. 2013), the medium resolution imaging spec-
trometer (MERIS) (Popp et al. 2011), the earth radiation budget experiment
(ERBE) (Li & Garand 1994), and the cloud albedo and radiation dataset from
AVHRR (CLARA-A1) (Karlsson et al. 2013). Among all the available satellite im-
ages, MODIS provides albedo data on a daily basis that is one of the most accu-
rate global land surface albedos at a 500-m spatial resolution and with a high
temporal resolution. Previous studies have shown that MODIS accurately pur-
veys land surface albedo from a radiometric and geolocation point of view
(Wolfe et al. 2002; Cescatti et al. 2012).

Objectives of the dissertation

This dissertation consists of two main parts. The first part, which was primarily
covered in Publications I and II, is dedicated to the monitoring of ecosystem
dynamics. The second part, which is linked to Publications III and IV, is dedi-
cated to the monitoring of albedo dynamics and its underlying mechanisms.
The main objectives of this dissertation are as follows:

• Objective I: To explore the potential of temporal and spatio-


temporal remote sensing data to generate early warning sig-
nals of critical transitions in wetlands and forest ecosystems

It was tested whether early warning signals of a critical transition can be de-
tected in unhealthy ecosystems with the known history of resilience loss in wet-
lands and forests. In addition, it was tested whether the new approach can im-
prove our ability to identify early warning signals of forest mortalities.

• Objective II: To investigate the effects of disturbances on al-


bedo dynamics in sparse forests in the Middle East

It was examined whether and how disturbances, including fires and droughts,
affect the temporal dynamics of albedo in the largest remaining forests in the
Middle East, i.e., the Zagros forests.

• Objective III: To assess the role of forest structure in modu-


lating albedo in forest ecosystems on a global scale

13
It was explored how forest structure, including forest density, tree cover, and
leaf area index, could explain shortwave, near-infrared, and visible albedo vari-
ations in forests around the world.

An overview of the publications

In the following, I briefly present the main outlines of each publication.

• In Publication I, the main focus was on exploring the state of a wetland eco-
system using temporal remotely sensed indices to verify the possibility of ac-
quiring early warning signals of a critical transition to rangeland. In addition, it
was shown that selecting the relevant key ecosystem variable, obtained from re-
mote sensing, contributes to the success of the approaches to identify early
warning signals of a critical transition. The state-of-the-art-knowledge reported
difficulties in predicting how close an ecosystem is to a critical transition due to
the lack of high-quality data and also in selecting appropriate ecosystem state
variables. A new remotely sensed composite index of vegetation and water,
called modified vegetation water ratio (MVWR), was proposed in this study. It
was shown that the new index has the potential to remarkably improve our abil-
ity to monitor the state of a wetland ecosystem. This publication reports the first
successful application of remote sensing data to identify a critical transition in
wetlands.

• Publication II, in line with Publication I, aimed to improve our ability to ex-
plore the dynamics of forests upon disturbances. In particular, a new approach
was developed for identifying the early warning signals of forest mortality, con-
sidering that the available methods are still subject to false alarming. In addi-
tion, a new R package, called “stew”, was developed. It enables a convenient as-
sessment of ecosystem dynamics using spatio-temporal data of satellite images
for users.

• In Publication III, the dynamics of land surface albedo upon disturbances


were investigated in the Middle East. The importance of this study comes back
to the major losses in vegetation cover in the studied region due to climate
change. In this publication, it was demonstrated how climate-induced disturb-
ances, such as severe droughts and fire events, influence albedo variations in
the Zagros mountains. This study presents the first report of the links between
albedo and disturbances in the eastern part of the Middle East (the Zagros
mountains).

• Publication IV presents the links between forest structure and albedo on a


global scale. Forests are particularly influential in regulating climate by altering
the earth’s surface albedo. In this publication, I examined the degree to which
forest structure can modulate albedo. This study provides the first map of the

14
Introduction

magnitude of the relationship between forest structure and albedo distributed


spatially across different pixels.

15
2. Literature review

Identifying ecosystem dynamics

2.1.1. Dynamic system theory

In the dynamic system theory, an important concept entitled “critical slowing


down” plays a key role in evaluating the system dynamics (Holling 1973). Based
on this theory, if there is a slowing down in the response of a system to gradual
effects of disturbance, it might lead to an abrupt change or a critical transition
in the state of the system (May 1977; Wissel 1984; Scheffer 2009). It means that
when a system is no longer resilient and cannot cope with external stresses, it
can consequently and abruptly respond to small perturbations that may ulti-
mately cause a state change (Holling 1973; Scheffer et al. 2012). As an example,
a critical transition in a forest may turn it towards an alternative state, including
a savanna, a rangeland, or a treeless state (Hirota et al. 2011).
According to dynamic system theory, by measuring the resilience of a system
close to a critical transition, we can evaluate the stability of a system (Holling
1973; Scheffer et al. 2001, 2012; Walters & Kitchell 2001). The reduced resili-
ence can be inferred as an early warning signal of a possible upcoming critical
transition (Scheffer 2010) or the signal of a system that is becoming unhealthy
(Dakos et al. 2015). Such characteristics can be specified via observing increases
in autocorrelation and high variability in the dynamics of a system (Dakos et al.
2012a). In some cases, the escalation of the skewness of the key system variables
might indicate an early warning signal of an upcoming critical transition
(Scheffer et al. 2012). The implication of this theory in the present study is to
measure the resilience of a system upon a disturbance and explore the degree to
which a system is approaching a critical transition (e.g., Scheffer et al. 2001;
Carpenter et al. 2011).
Applying the dynamic system theory to determining upcoming critical transi-
tions has found numerous applications in various scientific disciplines, ranging
from a social-ecological system (e.g., Nicholson et al. 2009) to climate (e.g.,
Dakos et al. 2008) and finance (e.g., May et al. 2008). However, the knowledge
gap for this goal is the applicability of these methods to monitor a complex sys-
tem, such as ecosystem dynamics. Such challenges limit the application of such
methods to mostly lab-scale (e.g., Drake & Griffen 2010), fully controlled exper-
iments (e.g., Carpenter et al. 2011), or simulation-based studies (e.g., Carpenter

16
Introduction

& Brock 2011). In the literature, at least two main concerns limit the applicabil-
ity of such methods to measure the dynamics of an ecosystem. First, it was
shown that the rate of success in getting a reliable early warning signal of a crit-
ical transition is dependent on the potential ecosystem state variable, i.e., the
variable that has the potential to present the state of an ecosystem (Dakos et al.
2015). Second, the methods to identify early warning signals are sensitive, still
subject to false alarms (Dakos et al. 2015; Nijp et al. 2019). The following sub-
chapter will focus primarily on these two major challenges.

Figure 2. A conceptual model that represents the dynamics of a system (e.g., ecosystem). A:
the state of a system that is highly resilient, and the potential of state change (y-axis) is low. In
this state, the rate of recovery from perturbations is relatively high. If disturbances push the state
of the system too far, represented by the ball in the figure, the ability of the system to shift the ball
back to the previous state (i.e., resilience) is high. B: the state of a system with low resilience. In
this state, even a small perturbation can dramatically change the state of the system. C: The
temporal signature of an ecosystem state variable from a healthy system. D: The temporal signa-
ture of an ecosystem state variable from an unhealthy system. Arrows indicate the parameters
that can identify resilience.

2.1.2. Ecosystem state variables

One important challenge limiting the application of dynamic system theory in


ecosystems is related to the selection of appropriate ecosystem state variables
for two main reasons. First, our knowledge about the choice of an ecosystem
state variable among a wide range of potential variables is still insufficient.

17
Some ecosystem state variables are not adequately sensitive to ecosystem dy-
namics and incorrectly show no alarm of an impending critical transition. Sec-
ond, an appropriate ecosystem state variable should have high-frequency obser-
vations in a way where it meets the requirements of the available methods
(Dakos et al. 2012a). The lack of such observations restricts the applicability of
these methods in identifying early warning signals of ecosystems. In practice, it
is difficult to measure the high-frequency time series of the right variables
through fieldwork, as discussed in subchapter 1.3. As an alternative, a multidi-
mensional remote sensing dataset can be exploited as an appropriate way to ac-
quire ecosystem state variables that make it possible to verify the early warning
signals of upcoming critical transitions.
To acquire high-frequency ecosystem state variables, various spectral indices,
derived from satellite images, provide a wealth of data needed for identifying
ecosystem dynamics. Different spectral indices can characterize different pro-
cesses and variables of an ecosystem over space and time. For example, the spec-
tral bands in which chlorophyll has the highest absorption (red band) and also
the highest reflectance regions (near-infrared band) purveys valuable infor-
mation about vegetation through the Normalized Difference Vegetation Index
(NDVI) (Tucker 1978). NDVI is one of the most frequently used indicators that
is linked to total vegetation cover. By monitoring NDVI, we can gain direct in-
formation about total vegetation cover and indirect information about several
biogeochemical cycles, e.g., the carbon cycle (Tucker 1978).
Although NDVI is reported to be adequately sensitive for the temporal analy-
sis of vegetation variations (e.g., DeVries et al. 2015), it becomes saturated at
high values of leaf area index (Sellers 1987; Zhang et al. 2000; Wang et al.
2004). Leaf area index (LAI) is a key structural characteristic of forest (both
understory and overstory). LAI is defined as the one-sided or hemisurface (half
of total) green leaf area per unit area of ground (Watson 1947; Chen & Black
1991). LAI has been applied for various purposes to recognize vegetation struc-
ture, such as characterizing vegetation changes upon environmental stress fac-
tors (e.g., Chen et al. 2005; Reygadas et al. 2019) and phenology changes
(Rautiainen et al. 2012). Other spectral bands, such as shortwave infrared
(SWIR) (1.4–2.5 μm), can be exploited to present leaf water contents (Woolley
1971), which, along with green band (0.55 to 0.57 μm), represents land surface
water areas, e.g., through the spectral index of Modified Normalized Difference
Water Index (MNDWI) (Xu 2006).

2.1.3. Methods to identify ecosystem dynamics

At least two groups of methods for exploring ecosystem dynamics have been
developed. The first group employs methods of temporal analysis to generate
early warning signals of a critical transition (see a comprehensive review of
approaches by Dakos et al. 2012b). The second group employs spatial metrics
over a spatial gradient, called space for time analysis, which is not the focus of

18
Introduction

this dissertation because it explores critical transition for a somewhat different


purpose (for more details see, e.g., Kéfi et al. 2014).
Temporal analyses of early warning signals measure the state of an ecosystem
in time through quantifying the variation in an ecosystem state variable aver-
aged over the whole ecosystem (Dakos et al. 2012a). The temporal analysis to
generate early warning signals (TEW) approach is typically achieved by statisti-
cal or mathematical evaluation of the variation of a state variable in an ecosys-
tem. It requires employing the statistical signature of, for example, autocorrela-
tion, and standard deviation. Such statistical methods, however, are sensitive to
pre-processing steps (e.g., selection of detrending methods or the choice of roll-
ing window size) in the TEW approach (TAMGA 2019). The reliability of the
results of the TEW approach can be considerably influenced when the size of the
data is not large enough. This is because almost 50% of the data is typically re-
served to estimate temporal analysis of early warning signals of a critical tran-
sition (Dakos et al. 2012b). There is a need to set a rolling window in the TEW
approach to capture the changes of statistical signature in a time series (e.g.,
autocorrelation). For ecosystems with quick changes, shorter rolling windows
can be appropriate, whereas ecosystems with slow changes need longer rolling
windows. To explore the dynamics of ecosystems, such as forests and wetlands,
a 50% rolling window is typically considered to provide a long enough rolling
window to measure the target statistical variable and short enough to provide a
sufficient number of observations to capture the changes (Dakos et al. 2012b).
Therefore, one main limitation in measuring ecosystem dynamics using tem-
poral analysis for generating early warning signals is the lack of a robust ap-
proach that is also not sensitive to preprocessing steps. Overcoming the afore-
mentioned limitations has been under slow progress in recent years (Angeler &
Allen 2016).

Albedo dynamics

The following subchapter describes the two main drivers of forest albedo, in-
cluding: 1) disturbances, such as droughts, fires, beetles, and fungi attacks, and
2) forest structure and forest type variations. My primary interest in this disser-
tation is black sky albedo (Fig. 3), defined as the ratio of the radiant flux for light
reflected from a unit surface area into the view hemisphere over the illumination
radiant flux when the surface is illuminated with a parallel beam of light from a
single direction (Schaepman-Strub et al. 2006).

19
Figure 3. A conceptual model that represents black sky albedo in a forest. According to this
model, the optical properties of overstory components (e.g., needle, leaf, and stems) and forest
floor components (e.g., snow, soil, and understory species) contribute to land surface albedo
values. Throughout this dissertation, the primary focus is on the confounding effects of the optical
properties of these components on black sky albedo, using moderate resolution imaging spectro-
radiometer (MODIS) satellite images.

2.2.1. Mechanisms underlying albedo variations in forests

Generally speaking, identifying albedo in forests is a challenging task. Part of


the challenge is related to the complexities of the radiation budget in a forest. In
a forest unit, the confounding effects of the optical properties of different com-
ponents, from leaf/needle to the spatial structure of overstory and forest floor,
can all affect the distribution of radiation, its pathways through a forest
(Stenberg et al. 2013; Kuusinen et al. 2014; Hovi et al. 2016), and consequently,
land surface albedo. Concerning the effects of each forest component, from nee-
dle or leaf to the spatial arrangements of forests, on reflectance: the reflectance
of a leaf/needle is low in the visible band where photosynthetic pigments are
actively absorbing solar radiation (Sims & Gamon 2002), while, on the other
hand, the reflection in the near-infrared band is high. SWIR absorption is as-
sumed to be larger than that of the near-infrared band and can further be raised
if water is available in the vegetation cover (Ceccato et al. 2001). Woody parts of
vegetation, such as stems, have large reflectivity in the visible and less reflectiv-
ity in the near-infrared bands compared with those parts of the vegetation cover
where the amount of photosynthetic pigments is high (Roberts et al. 2004). For-
est floors can also have considerably different reflectivity and optical properties
in snow-free conditions, depending on soil types and understory species (e.g.,
Rautiainen et al. 2011; Rautiainen & Lukeš 2015). In addition, different leaf
types (broadleaf and needleleaf) and phenology (evergreen and deciduous) can
considerably influence albedo. For example, an evergreen needleleaf tree can
have a higher multiscattering albedo than broadleaved trees. It can enhance the

20
Introduction

likelihood of photon recollision (Lukeš et al. 2013b) and, in turn, make albedo
decrease in the forest landscape.
The measurement of the abovementioned parameters in forests has been stud-
ied in a number of works. For example, Lukeš et al. (2013b) and Hovi et al.
(2017) studied directional-hemispherical reflectance and transmittance factors,
such as the optical properties of the needle/leaf. They showed that tree species
are different in shortwave-infrared, e.g., in pine and spruce tree species, the
transmittance of needles is considerably lower than their reflectance. Other
studies demonstrated that forest floor (Kuusinen et al. 2014; Rautiainen &
Lukeš 2015) and overstory (Kuusinen et al. 2012) could have significantly dif-
ferent optical properties and albedo in forests. It has been shown that an in-
crease in stand biomass and canopy cover can result in the decrease of albedo
in a forest landscape (Lukeš et al. 2013a).
Although numerous works have studied the mechanisms controlling albedo in
forests, the majority of the previous studies have been performed only on local
to regional level and mainly in the Boreal region (e.g., Betts & Ball 1997; Betts
2000; Rautiainen et al. 2011; Lukeš et al. 2013a, 2014; Bright et al. 2014, 2018;
Kuusinen et al. 2016; Hovi et al. 2019). This suggests a clear geographical gap
in the understanding of albedo dynamics. Such geographical gaps may not only
result in the inefficiency of climate mitigation plans on a global scale but can
also produce contradictory results. For example, the contradictory findings of
Sun et al. (2010) compared with Dore et al. (2012) or Lukeš et al. (2013a) could
be attributed to the inconsistency of the optical properties of forest components
in each study site.
To avoid this shortcoming and to adequately evaluate albedo dynamics, a sen-
sitivity analysis for the physical models of land surface albedo can be exploited
(Wang 2005) to evaluate how albedo is changing in response to the changes in
forests. It should also be noted that the current physical models of albedo sim-
ulation extensively rely on a spectral library of soils, leaves, needles, and other
components of forests (Kuusk & Nilson 2000). Additionally, different sources
of uncertainties, such as model inputs as well as model formulation and as-
sumption, affect the modeling of forest reflectance (Hadi 2018). Finally, oper-
ating such physical models pixelwise on a global scale could be computationally
expensive. Bridging this knowledge gap in the literature using statistical meth-
ods has been one of the main objectives of the present study (see subchapter
2.2.3).

2.2.2. Forest albedo variations upon disturbances

Investigating the temporal dynamics of albedo under the influence of disturb-


ances (e.g., droughts or fires) has been the subject of numerous studies (e.g.,
Govaerts & Lattanzio 2008; Lyons et al. 2008; Lee et al. 2011; O’Halloran et al.
2012; Gatebe et al. 2014; Vanderhoof et al. 2014). However, the reported obser-
vations have sometimes been controversial. For example, a number of works
have claimed that forest loss due to disturbances has led to an albedo decrease

21
(e.g., Veraverbeke 2010) while other works have reported the opposite observa-
tions (Govaerts & Lattanzio 2008). Some potential reasons for the mentioned
inconsistencies are outlined here.
The variations in forest structure components can affect reflectance and, in
turn, albedo (subchapter 2.2.1). The disturbance of changing forest structure
can further alter forest albedo (Schwaiger & Bird 2010). Furthermore, the type
of disturbance can also affect albedo in a different manner (O’Halloran et al.
2012; Dintwe et al. 2017).
For example, forest albedo may decrease under the influence of fires because
black carbon in the soil and on the boles of dead trees can reduce albedo by
optically darkening the background (Chambers & Chapin 2002). When a fire is
the main disturbance, the establishment and growth of optically bright herba-
ceous plants and shrubs can rapidly increase albedo after the fire occurrence
(Lyons et al. 2008). In addition, pre-disturbance vegetation structure, soil prop-
erties, vegetation regrowth, recovery rates, and time and duration of fire dis-
turbances can all contribute to albedo alterations (Gitas et al. 2012).
In contrast, drought conditions can increase albedo through the loss of over-
story and greater exposure of optically brighter snow‐covered surfaces or bare
soil (Charney 1975; Govaerts & Lattanzio 2008). The frequency and intensity of
the drought effects and albedo feedbacks all depend on multiple factors, such as
forest structure and composition (subchapter 2.2.1, Schwaiger & Bird 2010;
Gitas et al. 2012), the severity of drought (i.e., lack of precipitation), and other
conditions, such as topography and relative humidity.
It, however, is still poorly understood how and to what extent the magnitude
of albedo changes under the influence of forest disturbances. To explore the im-
pacts of climate-induced disturbances on ecosystems, physical climate models
already shed some light on the processes and mechanisms (Pitman et al. 2009).
Nevertheless, the usage of such physical climate models is still limited due to
their coarse spatial resolution, uncertainties in physical processes, parameteri-
zations of albedo, and input data (Pitman et al. 2009; Li et al. 2015). Therefore,
this dissertation primarily focused on statistical analysis to address the research
questions (see subchapter 2.2.3).

2.2.3. Statistics to measure the dynamics of albedo

To monitor the changes of the response variable (e.g., albedo), various statistical
methods can be exploited, such as thresholding (e.g., Hayes & Sader 2001), dif-
ferencing (e.g., Zheng & Basher 1999), segmentation (e.g., Chance et al. 2016),
statistical boundary (e.g., Xue et al. 2014), and regression (e.g., Sonnenschein
et al. 2011). For example, the thresholding method measures the variations in
an ecosystem using a time series of target variables. Observation of a change
that exceeds the predefined threshold is considered a change in this method. In
the segmentation method, the time series of a target variable is segmented into
a series of straight lines using the residual-error and angle criteria. Deviation in
the resulting straight-line segments at the pixel level is then marked as abrupt

22
Introduction

or gradual changes. In the statistical boundary method, the time series of the
target variable is analyzed following a predefined statistical boundary. If a sig-
nificant deviation is detected in the time series of a target variable, this method
considers it a change.
Among all the available methods, break detection for additive season and
trend (BFAST) (Verbesselt et al. 2012a) is among the most widely used ones for
change detection in a time series following a disturbance occurrence (Zhu 2017).
It decomposes the time series of a target variable into the trend, season, and
noise components. If a significant change makes the segments of the time series
deviate from the statistical boundary, it will be treated as either an abrupt or
gradual change. The successful application of the BFAST method has been ver-
ified for the monitoring of forest changes (Verbesselt & Hyndman 2010), varia-
tions in seasonality (Verbesselt et al. 2010), and near real-time detection
(Verbesselt et al. 2012a). Among regression-based methods, Theil-Sen’s slope
(Sen 1968), as a nonparametric test, is widely accepted as a robust method to
identify outliers and can be used to measure the magnitude of change in the
response variable over time. The differencing method can be used when report-
ing the exact magnitude of change in the response variable is desirable.

23
3. Materials

Description of study sites

To benchmark the developed methods presented throughout this dissertation,


several study sites at various scales, taken from a limited geographical location
to a global-scale study, were investigated (Fig. 4).
The possibility of identifying early warning signals of critical transition was
investigated in Publication I via studying two wetlands, one healthy and the
other unhealthy. Both of the wetlands are designated as Ramsar Sites (conven-
tion on wetlands of international importance) and are considered as important
hosts for migratory waterbirds1, which highlights the importance of these study
sites. The studied unhealthy wetland, called Dorge Sangi, located in the Urmia
basin, Iran, has suffered from a severe drought in the past decades
(Hassanzadeh et al. 2012). In the context of this dissertation, drought refers to
“prolonged absence or marked deficiency of precipitation” (IPCC 2007). The
drought, alongside anthropogenic effects, has caused a severe decrease in the
water content in Dorge Sangi, resulting in an increase in water salinities on one
hand and a critical transition in the vegetation community on the other hand
(Ahmadi et al. 2011; Shadkam et al. 2016). The studied healthy wetland was the
Lake Arpi located in a national park in Armenia. According to my sources, there
are no reports on major disturbances that could cause a resilience reduction in
the Arpi wetland for the period between 2000 and 2016. I selected the Arpi wet-
land as an evidence study site because it has a reasonably similar system to
Dorge Sangi from an eco-hydrological point of view. The remarkable similarities
between the two wetlands stem from them both being freshwater, relying on
precipitation to feed, and having the same mean annual temperature (12.1°C in
Dorge Sangi and 11°C in Arpi) and the same consistent mean annual precipita-
tion (471 mm in Dorge Sangi and 550 mm in Arpi).
To investigate the dynamics of forests upon disturbances in Publication II, I
considered three study sites. The study sites included two unhealthy forest eco-
systems (Sequoia and Yosemite located in national parks in America) which are
no longer neither stable nor efficiently resilient (see Liu et al. 2019), as well as
one healthy study site in Canada (Wood Buffalo national park). According to my
sources, no sudden forest mortality has been reported in the literature for these

1
https://rsis.ramsar.org/ris/45, accessed date: 18.05.2020

24
Materials

study sites. As all the selected study sites are located in national parks, it is worth
mentioning that no exploitation of natural resources is expected for them.
To study the effects of disturbances such as fungi attacks and drought-induced
climate change on forest albedo in Publication III, I considered four study sites
in the Middle East. In that region, the evidence of climate-induced disturbance
effects in forests due to drought, fire, fungus, and beetle attacks have become a
major concern (Allen et al. 2010; Hoerling et al. 2012; FAO 2020). The drought
conditions have resulted in temperature anomalies up to 7℃ in the Middle East
since 1960 (Kaniewski et al. 2012; Lelieveld et al. 2012), which have affected
ecosystems in the region, especially those in Iran (Voss et al. 2013). Also, several
studies have illustrated a consistent climate change over the region in the last
decades (Kaniewski et al. 2012; Lelieveld et al. 2016). Such climate-related
changes have resulted in extensive oak tree mortality in the Mediterranean ba-
sin since the 1980s (Clara 2013). The drought-induced climate change followed
by fungus and beetle attacks has also been observed in forests in Iran (Hosseini
et al. 2012) and Turkey (Jurc & Ogris 2006). The Middle East has been recog-
nized as one of the global hot spot regions that are expected to be faced with
severe climate changes in the near future (Giorgi & Lionello 2008; Lelieveld et
al. 2012). On top of making the lives of indigenous people difficult, drought-
induced disturbances increase the frequency of forest disturbances in the few
remaining areas of tree cover in the area (Barlow et al. 2016). In Publication III,
the main focus was on forests in the entire Zagros region as well as four study
sites in the same region, namely Marivan, Ilam, Ize, and Dashte baram. The se-
lected sites represent different types of forest disturbances, e.g., fires in Ilam
and Ize and droughts in Marivan and Dashte baram, which have resulted in var-
ious degrees of tree defoliation. The Ilam, Ize, Marivan, and Dashte baram study
sites also suffered from fungus and beetle attacks.
The links between forest structure (forest density, tree cover, and LAI) and
land surface albedo, as the main aim of Publication IV, were investigated in pure
forests. A pure forest pixel is defined as forests where all the 500-m spatial res-
olution pixels have the same forest type within a 1-km spatial resolution pixel
on a global scale (more information in Publication IV). In this study, I employed
the high-quality pure forest pixels based on MODIS land cover product, cover-
ing a total of 7% of the earth’s surface. The forests were grouped into 20 classes,
namely sub-biomes, by using the ecoregions map to obtain major biome-zones
(Olson et al. 2001) and a landcover map to identify forest types (Friedl & Sulla-
Menashe 2019). The major biome-zones based on the ecoregion map included
boreal, mediterranean, temperate, and tropical regions. The forest types were
also categorized based on the land cover map into evergreen needleleaf (EN),
evergreen broadleaf (EB), deciduous needleleaf (DN), deciduous broadleaf
(DB), mixed forests (Mixed), and woody savannah forests (WS).

25
Figure 4. The location of the study sites in this dissertation. The blue circles represent the lo-
cation of wetlands in Publication I. The green circles represent the location of national forest
parks in Publication II. The red box represents the Middle East, and the four red circles repre-
sent study sites chosen for Publication III. The grey lines represent the global scale that was the
boundary of the study in Publication IV.

Data acquisition

All resources used in the carried-out research were provided by the freely avail-
able remotely sensed data (for details see Table 1), except for the high spatial
resolution satellite images such as IKONOS, Pléiades, and Quickbird for tree
cover estimations in Publication III. A major part of the applied resources was
obtained from MODIS data, accompanied by other satellite-based products re-
trieved from Landsat and radar data.

26
Materials

Table 1. Satellite-based data retrievals were used in this dissertation. The full list can be found
in each publication.

Name Unit Product Spatial Year Publication Data


resolution and source
sensor
Reflectance
reflectance - MOD09A1 500-m 2001- I and III Vermote
(MODIS) 2014 2015
albedo - MCD43A3 500-m 2000- III Schaaf and
(MODIS) 2016 Wang 2015

BRDF1 - MCD43A1 500-m 2000- IV Schaaf and


parameters (MODIS) 2016 Wang 2015

Climate variables and other data


precipitation mm/hr CESM1- 0.5° 1950- I and III Hurrell et
BGC, and (multiple 2016 al. 2013;
NCEP/NCAR2sensors) Kalnay et al.
1996
FIRMS3 - MCD14DL 500-m 2000- III FIRMS
(MODIS) 2016 2018
ecoregions - - - 1973- IV Olson et al.
2000 2001
Vegetation cover
forest number - 1-km 2005 IV Crowther
density of (multiple et al. 2015
trees/𝑘𝑚 2 sensors)
tree cover % VCF4 30-m 2003- IV Sexton et
(Landsat) 2008 al. 2013
forest type - MCD12Q1 500-m 2005 II and IV Friedl and
(MODIS) Sulla-
Menashe
2015
leaf area m2/m2 MCD15A2 500-m 2000- III and IV Knyazikhin
index (MODIS) 2016 et al. 1998
normalize - MOD13A1 500-m (MODIS) 2000- I and II Didan
difference 2019 2015
vegetation
index
Ancillary data
BRDF- - MCD43A2 500-m 2005 III and IV Schaaf and
albedo (MODIS) Wang 2015
quality
solar zenith degree MCD43A2 500-m 2005 IV Schaaf and
angle (MODIS) Wang 2015
aerosol MCD43A2 500-m 2005 IV Schaaf and
optical depth (MODIS) Wang 2015
snow - MCD43A2 500-m 2005 III and IV Schaaf and
(MODIS) Wang 2015
digital m - 90-m - III and IV Jarvis et
elevation (SRTM5) al. 2008
model
water area - JRC6 30-m 2005 IV Pekel et al.
(Landsat) 2016
1 bidirectional reflectance distribution function

27
2 community earth system model and National centers for environmental prediction/national
center for atmospheric research
3 fire information for resource management system
4 vegetation continuous fields (VCF)
5 shuttle radar topography mission (SRTM)
6 Joint Research Center (JRC)

Data extraction and quality assessment

For all carried-out research, the data extractions (i.e., extracting data from dif-
ferent pixels within the geographical location of a study site) and aggregations
(e.g., aggregating the resolution from 500-m to 1-km in Publication III) were
implemented through the mean function. The only exception was the land cover
map, which was categorical data and for which a modal function was applied. In
all the publications, as standard procedure, I used only high-quality satellite
data, using the ancillary information of each product.
For albedo studies, further quality assessments were carried out on top of
standard procedures. Those involved using high-resolution slope data (i.e., 90-
m spatial resolution) to avoid the error in forest albedo estimates located in high
slope locations due to topography-related issues by excluding pixels with high
slope values (slope >10%). For Publication IV, high-resolution water data (i.e.,
30-m spatial resolution) was employed to avoid very low reflectance of water
that might cause uncertainties in albedo values (water >5%).
Depending on the requirements of individual studies, different approaches
were implemented to treat missing satellite data. For example, in Publications I
and III, the selected sites were in the Middle East where the ecosystems had
suffered from drought conditions for a long time. As a result, the number of
high-quality data was large enough during the studied period. Therefore, it suf-
ficed to simply remove the missing values from the analyses and fill in the miss-
ing ones using nonlinear models.
One of the study sites considered in Publication II (Wood Buffalo national
park in Canada) was located in the boreal zone, characterized by long winters,
typically cloudy skies and high solar zenith angles in satellite observations. For
this study site, after data quality assessments, I noticed that the number of miss-
ing values was considerably large during the wintertime. For this case, although
some good quality pixels were sparsely available in the wintertime, I excluded
the wintertime observations (from December to March) instead of filling the
missing values or removing single low-quality observations.
For Publication IV, which was a global-scale study, the main focus was on the
peak growing season to avoid the effects of a varying number of missing values
throughout the year in the analyses. The missing satellite data due to low-quality
observations during peak growing season was then removed from the datasets.

28
Methods

4. Methods

Overview of the applied methods and tools

All the computations and data analyses carried out in this dissertation were per-
formed using freely available open-source software and platforms, including R
statistical software (R Core Team 2012), QGIS (Team 1991), and Google Earth
Engine (Gorelick et al. 2017). Details of the specific application of each of the
mentioned tools are reported in Table 2.
In the following, a summary of all methods employed in the present work is
presented.

29
Table 2. A list of main statistical analyses and software used in this dissertation. A full list can
be found in each publication.

Publication Objective Statistical Data Software


analyses type
I I Autocorrelation, Temporal R statistical software (R
variance, and Core Team, 2013). R
skewness in TEW1 packages including
approach “earlywarnings”(Dakos et al.
2012a), and “raster” (for
fundamental raster analysis)
(Hijmans & van Etten 2012).

II I SPT2, STM3 and Spatio- R statistical software (R


STG4 approaches temporal of Core Team 2013), QGIS
data (QGIS Development Team,
1991), Google Earth Engine
(Gorelick et al. 2017), and my
new R package, called “stw”.

III II Breaks for Temporal R statistical software (R


additive seasonal Core Team, 2013), R
and trend method, packages including
Pearson “bfast”(Verbesselt &
correlation Hyndman 2010), and “raster”
(for fundamental raster
analysis) (Hijmans & van
Etten 2012).

IV III Generalized Spatial R statistical software (R


additive model, Core Team 2013), QGIS
Sen’s slope, and (QGIS Development Team
Pearson 1991), and Google Earth
correlation Engine (Gorelick et al. 2017).
R packages, including
“raster” (for fundamental
raster analysis) (Hijmans &
van Etten 2012), “rts” (for
time series analysis) (Naimi
2016), “mgcv” (for GAM5)
(Wood 2011), “phenopix” (for
detecting peak growing
season) (Filippa et al. 2016),
“usdm” (for multicollinearity
test) (Naimi et al. 2014),
“ggplot2” (Wickham 2016),
and “rasterVis” (for
visualizations) (Lamigueiro &
Hijmans 2018).

1 temporal analysis of early warning signals


2 spatial patterns of temporal analysis
3 spatio-temporal analysis of Moran’s 𝐼

4 spatio-temporal analysis of Geary’ 𝑐

5 generalized additive model

30
Methods

Monitoring ecosystem dynamics

Throughout this thesis, three approaches were used to identify the ecosystem
dynamics using early warning signals of critical transition methods in wetlands
or forests. The first approach, known as the temporal analysis generating early
warning signal (TEW) approach, employs a time series, which is extracted from
a study site, to test whether the early warning signals can be obtained (more
details in subchapter 4.2.1). The TEW approach was the main applied method
for Publication I for monitoring the state of the wetland ecosystems.
The second approach, namely the spatial pattern of temporal analysis (SPT),
applies the TEW approach to each pixel to generate a spatial pattern of early
warning signals (more details in subchapter 4.2.2). The third approach, which
is the spatio-temporal analysis of early warning signals based on Moran’s 𝐼
(STM) and Geary’s 𝑐 method (STG), estimates the trend of changes in the local
spatial autocorrelation statistics over time (more details in subchapter 4.2.3).
The TEW and SPT approaches, in addition to the proposed approaches in this
dissertation, STG and STM, are used to detect early warning signals of forest
mortality in Publication II. A conceptual diagram representing the differences
among the abovementioned approaches (TEW, SPT, STM, and STG) can be seen
in Fig. 1 of Publication II.

4.2.1. Temporal early warning signals (TEW)

To carry out the analysis of the TEW approach, I first estimated the monthly
spectral indices of NDVI, MNDWI, and modified vegetation water ratio
(MVWR) as the main considered ecosystem state variables in healthy and un-
healthy wetlands (study sites are introduced in subchapter 3.1) between 2001
and 2014.
The NDVI index is reported to be an appropriate index for monitoring vegeta-
tion dynamics in wetlands (e.g., Petus et al. 2013) and is calculated via (Solano
et al. 2010):

𝐵 𝑛𝑖𝑟 – 𝐵𝑟𝑒𝑑
NDVI =
𝐵 𝑛𝑖𝑟 +𝐵𝑟𝑒𝑑
(1)

where 𝐵𝑟𝑒𝑑 and 𝐵 𝑛𝑖𝑟 are the MODIS surface reflectance factors in the red band
(band 1, 0.620 to 0.670 μm) and the near-infrared band (band 2, 0.841 to 0.876
μm), respectively.
The MNDWI is shown to be one of the most accurate spectral indices in deter-
mining the water area (Rokni et al. 2014) and is calculated as follows:

𝐵 –𝐵
MNDWI = 𝐵 𝐺𝑟𝑒𝑒𝑛 +𝐵 𝑆𝑊𝐼𝑅 (2)
𝐺𝑟𝑒𝑒𝑛 𝑆𝑊𝐼𝑅

where 𝐵 𝑆𝑊𝐼𝑅 refers to the surface reflectance of the SWIR MODIS band (band
6, 1.23 to 1.25 μm), and 𝐵 𝐺𝑟𝑒𝑒𝑛 refers to the green MODIS band (band 4, 0.55

31
to 0.57 μm). The efficiency of MNDWI in monitoring the wetland ecosystem has
been verified in a number of studies (see, e.g., Xu 2006).
MVWR is a new index developed and presented for the first time in Publica-
tion I. The MVWR is a modified version of the vegetation water ratio (VWR)
index, which describes the relationship between vegetation and water spectral
indices (Zhao et al. 2009). The new index (i.e., MVWR) seems to be a more ap-
propriate representative of the wetland ecosystem compared to the VWR index
because the NDVI and MNDWI were rescaled to a range between 0 and 1 to
avoid the effect of negative values in the calculation of MVWR. Then, the ln
function was used to produce a logarithmic scale of the results and show the
range of the results in a compact way. More information about the weaknesses
of VWR and how MVWR modified those weaknesses can be found in Publication
I and the Discussion chapter.

NDVI+1
MVWR = 𝑙𝑛 (MNDWI+1) (4)

Following the calculation of ecosystem state variables in each study site, their
respective time series were extracted. As the next step, the time series were
detrended to eliminate the effects of non-stationary behaviors using detrending
methods (e.g., first differencing) available in the “earlywarnings” R package,
and the robustness of the various detrending methods was evaluated using the
Augmented Dickey-Fuller test. After eliminating the effects of the non-station-
ary condition, a rolling window of 50% was selected to measure the TEW of for-
est mortality based on a sensitivity analysis (see Publication I for details).
To explore the state of healthy and unhealthy wetland ecosystems (Publication
I), I tested the temporal statistical signatures of the autocorrelation, standard
deviation, and skewness parameters in each ecosystem state variable. An in-
creasing trend in the aforementioned statistical signatures could be considered
an indicator of early warning signals of a critical transition. Accordingly, the
trend of changes in statistical signatures was characterized using Kendall’s τ,
which is a nonparametric approach to statistically describe the monotonic up-
ward or downward trends in a target variable in time (Kendall 1948). Kendall’s
τ of zero means no association in the trend of change, while the values of +1.0
or -1.0 mean perfect increasing and decreasing trends, respectively.
The autocorrelation function (ACF) at lag-1 is calculated as:

𝐸[(𝑧𝑡 −𝜇)(𝑧𝑡+1 −𝜇)]


ACF = 𝛿𝑧2
(5)

where 𝐸 is the expected value, 𝜇 is the mean, and 𝛿 is the variance of the variable
𝑧. 𝑧𝑡 and 𝑧𝑡+1 are the values of the variable 𝑧 obtained at time 𝑡 and 𝑡 + 1 (lag-
1), respectively.
Standard deviation (SD) is the second moment around the mean of a distribu-
tion and can be expressed as a standard deviation via:

32
Methods

1
SD = √𝑛−1 ∑𝑛𝑡=1(𝑧𝑡 − 𝜇)2 (6)

where 𝑛 refers to the number of records.


Skewness (SK) can also serve as a leading indicator of upcoming abrupt
change and is calculated via:

1 𝑛
∑ (𝑧 −𝜇)3
𝑛 𝑡=1 𝑡
SK = 1 𝑛
(7)
√ ∑𝑡=1(𝑧𝑡 −𝜇)2
𝑛

4.2.2. SPT analysis in forests

I used the NDVI index to explore the state of healthy and unhealthy forests
(study sites are introduced in subchapter 3.1) between 2000 and two years prior
to the abrupt change at each study site, using an analysis based on the TEW
approach (results are available in Publication II), and the SPT, STM, and STG
methods (corresponding results can be seen in Fig. 7). The timing of the abrupt
change was estimated using the BFAST method between 2000 and 2019, as
elaborated in Publication II.
The NDVI has shown to be an appropriate ecosystem state variable in repre-
senting the dynamics of forests (Rogers et al. 2018; Liu et al. 2019) where the
LAI is not very high, as mentioned in subchapter 2.1.2. It should be noted that
water-based indicators (e.g., the normalized difference water index, NDWI)
(Gao 1996) is reported to be a more sensitive indicator in monitoring evergreen
forests (Delbart et al. 2005; Wu et al. 2014). Nevertheless, the main reason for
considering the vegetation-based indicator (i.e., NDVI) in Publication II was the
presence of fungi attacks as one of the disturbances (Liu et al. 2019). It has been
shown that NDWI is not sensitive enough to identify the effects of diseases and
fungi attacks in forest variations (Vescovi 2017). The corresponding signals of
NDWI were shown in the supplementary file in Publication II.
In the SPT approach, the TEW approach was applied (subchapter 4.2.1) on all
high-quality pixels of the monthly NDVI time series within the geographical ex-
tent of each study site. I used the first differencing method to detrend the time
series following the Augmented Dickey-Fuller test. I applied a rolling window of
50% on the time series to estimate temporal autocorrelation at lag 1. Next, Ken-
dall’s τ values, calculated for both the autocorrelation of NDVI and the time se-
ries of NDVI values, were assigned to each high-quality forest pixel (correspond-
ing results can be seen in Fig. 6–7).
Although both SPT and TEW analyses are sensitive to preprocessing (e.g.,
detrending), the SPT analysis is expected to be a more efficient method in iden-
tifying the state of ecosystems compared to the TEW approach. The reason is
that it allows generating the spatial distribution of signals and produces multi-
ple observations from different pixels, whereas the TEW approach generates a
single value as a general signal of the whole ecosystem.

33
4.2.3. Spatio-temporal analyses of early warning signals

To implement the STM and STG methods, two local spatial autocorrelation sta-
tistics, namely local Moran’s 𝐼 and local Geary’s 𝑐, were first applied to the high-
quality monthly NDVI values at each time step, between 2000 and two years
before abrupt change occurrence for each study site (for more details, see Pub-
lication II). Declines in local spatial autocorrelation infer a high degree of forest
fragmentations, meaning that a forest is increasingly divided into separated
portions with smaller and more isolated patches. In such situations, a critical
transition is expected to impend (Taubert et al. 2018). Therefore, observing de-
clines in local spatial autocorrelations serve as early warning signals of critical
transition. Local Moran’s 𝐼 at site 𝑖 was calculated via (Anselin 1995):

(𝑥𝑖 −𝑥)
𝐼𝑖 = 𝑠2
∑𝑗 𝑤𝑖𝑗 (𝑥𝑗 − 𝑥) (8)

where, 𝐼𝑖 is the local Moran’s 𝐼 for site 𝑖, 𝑤𝑖𝑗 is the spatial weight for the obser-
vation 𝑗 (often binary, i.e., 1 for neighboring locations and 0 elsewhere), 𝑥 and
𝑠 2 are the global sample mean and variance of the observations, respectively,
and 𝑥𝑖 and 𝑥𝑗 are the observed values at neighboring sites 𝑖 and 𝑗, respectively.
Positive values of 𝐼𝑖 indicate a cluster of similar values around site 𝑖, while neg-
ative values imply dissimilarity between the site 𝑖 and its neighboring areas.
Local Geary’s 𝑐 statistic at site 𝑖 was calculated as (Anselin 1995):

1 2
𝑐𝑖 = 𝑠2 ∑𝑗 𝑤𝑖𝑗 (𝑥𝑖 − 𝑥𝑗 ) (9)

with the same variables as local Moran’s 𝐼. This statistical tool quantifies a
standardized squared distance between the values at site 𝑖 and its neighboring
areas. High values of 𝑐𝑖 indicate substantial diversity between site 𝑖 and its
neighbors, while low values imply a similarity between them. As a comparison,
local Moran’s 𝐼 measures deviations of locations from the global mean, while
local Geary’s 𝑐 represents differences between a site and the values of its neigh-
boring sites.
As the next step, the values of the local spatial autocorrelation at each pixel
between 2000 and two years prior to the abrupt change were extracted. The ex-
tracted values were then smoothed using a moving average with the size of three
(i.e., three months) to decrease the effects of potential errors, such as those may
help in preventing the risk of the false alarm due to precipitation occurrence,
which can increase autocorrelation. Finally, the trend of changes in the time se-
ries of local spatial autocorrelation was calculated at each pixel using Kendall’ τ
test (corresponding results can be seen in Fig. 7).

34
Methods

Albedo dynamics

4.3.1. Black sky albedo

In this dissertation, I used snow-free black sky albedo (a MCD43A3 product)


because it mirrors only land surface albedo information without disrupting fac-
tors such as snow and atmospheric conditions (Strahler et al. 1999). It allows
for isolating the influence of varying factors to forest structure only to explore
albedo variations (Publications III and IV).
The MCD43A3 is a readily available albedo product that corresponds to local
noon illumination conditions. In the MCD43A3 product, albedo is estimated
through the following steps: 1) performing atmospheric correction and convert-
ing the top of atmosphere radiance to surface directional reflectance; 2) con-
verting directional reflectance to spectral albedo using the bidirectional reflec-
tance distribution function (BRDF) angular conversion; 3) converting spectral
albedos to broadband albedos using a narrowband-broadband conversion algo-
rithm (Schaaf et al. 2002; Liang 2005). BRDF is defined as the scattering of a
parallel beam of incident light from one direction in the hemisphere into an-
other direction in the hemisphere (Schaepman-Strub et al. 2006).
The solar zenith angle (SZA), which is defined as the angle between the zenith
and the sun position, is a parameter that can affect the reflectance and, conse-
quently, the albedo (Hovi et al. 2016; Ni & Woodcock 2000). For example, it is
found that the proportion of diffuse radiation can increase by increasing SZA
(Kuusinen 2014). The higher the SZA, the higher the path length and radiation
attenuations. It results in a decrease in the direct beam radiations that may
reach the earth’s surface (Kuusinen 2014). It should be noted that for a global
study, eliminating the effects of SZA variations might have a crucial role in al-
bedo studies (e.g., Publication IV). To this end, I estimated how and if SZA var-
iations may influence albedo during the peak growing season on a global scale
in a separate publication (Alibakhshi et al. 2020a). I calculated the albedo at the
fixed solar zenith angles of 28° and 38° and compared the results with albedo
values corresponding to the local solar noon (derived from MCD43A3), which
ranged from SZA of 0° to 68° (see Alibakhshi et al. 2020a and Table A3 of
Publication IV). The results showed that albedo at fixed SZA was similar to al-
bedo values corresponding to the local noon SZA during peak growing season
on a global scale (Alibakhshi et al. 2020a). This phenomenon was opposite to
the effects of the diurnal variation of albedo with different values of SZA (Lukeš
et al. 2013a).
Accordingly, the albedo was estimated at the fixed solar zenith angle of 38° at
shortwave (300–5000 nm), near-infrared (700–5000 nm), and visible (300–
700 nm) bands in Publication IV. The SZA and BRDF were used to calculate
black sky albedo at fixed SZA, evaluated at three different spectral regions via
(Strahler et al. 1999):

35
𝛼𝑏𝑠 (𝜃, 𝜆) = 𝑓𝑖𝑠𝑜 (𝜆) + 𝑓𝑣𝑜𝑙 (𝜆)(−0.007574 − 0.070987 × 𝜃 2 + 0.307588 × 𝜃 3 ) +
𝑓𝑔𝑒𝑜 (𝜆)(−1.284909 − 0.166314 × 𝜃 2 + 0.041840 × 𝜃 3 )
(10)

where 𝛼𝑏𝑠 (𝜃, 𝜆) refers to the black sky albedo and describes the albedo under
direct illumination conditions ignoring diffuse illuminations at band λ, and 𝜃 is
the SZA of 38°. The 𝑓𝑖𝑠𝑜 (𝜆), 𝑓𝑣𝑜𝑙 (𝜆), and 𝑓𝑔𝑒𝑜 (𝜆) refer to the three weighting
parameters in the model (isotropic, volumetric, and geometric, respectively) for
λ (i.e., SW, NIR, and VIS).

4.3.2. The temporal dynamics of albedo

To study the relationships between albedo and disturbances, two sets of anal-
yses were carried out in the Zagros forest between 2000 and 2016. First, I ex-
plored how precipitation rate, the number of fire events, burn severity, and LAI
explained albedo in the entire Zagros forest using the Pearson correlation test
(subchapter 3.1). Second, I analyzed how albedo was changing in response to
disturbances, particularly fire and drought, using the BFAST method in four
study sites in the Middle East (subchapter 3.1). I also quantified the magnitude
of climate change in the Middle East, and I explored the dynamics of changes in
LAI and tree cover to understand how disturbances affected understory and
overstory (results in Publication III).
To explore the temporal dynamics of albedo, the forest pixels within the north-
ern and southern Zagros areas, as well as in the four study sites, were first de-
lineated. Next, the high-quality monthly time series of the response variable
(i.e., albedo), as well as predictors (i.e., precipitation rate, the number of fire
events, burn severity, and LAI), were obtained or estimated between 2000 and
2016 (see Table 1 and Publication III for details).
The normalized burn ratio (NBR) and the difference NBR (dNBR) were esti-
mated using near-infrared and shortwave infrared bands, derived from the
MOD09A1 version 6 MODIS product. Accordingly, the NBR index was calcu-
lated via (Key & Benson 2006):

NIR−SWIR
NBR = (11)
NIR+SWIR

Furthermore, dNBR was quantified by applying the bi-temporal image differ-


encing on pre- and post-fire NBR data (one year) as follows (Key & Benson
2006):

dNBR = NBR pre−fire − NBR post−fire (12)

To define the pre- and post-fire data, I first considered only the summertime
(June to August), as suggested by Veraverbeke et al. (2010b). The NBR data
from the summer before and after the fire event were then aggregated to obtain

36
Methods

the pre-fire values and post-fire values, respectively. More details on this ap-
proach can be found in Publication III.
I then used a Pearson correlation test to evaluate the relationship between the
predictors and response variables in the entire Zagros forest area (correspond-
ing results can be seen in Fig. 8 and Tables 2–3 of Publication III).
For studying the albedo dynamics at the four sites, I calculated a possible
breakpoint that might happen in the time series within a specific window using
the BFAST method (Verbesselt et al. 2012b). In here, a breakpoint represents a
significant change in the time series of a response variable that is employed to
represent the effects of an identifiable forest disturbance on albedo.
To estimate a possible breakpoint in the time series of data, decomposition of
the time series of data into the trend and seasonal components was carried out
as the first step. To do so, a piecewise linear trend and a seasonal model were
fitted to the data (from time points 1 to n) by (Verbesselt & Hyndman 2010):

𝑌𝑡 = 𝑇𝑡 + 𝑆𝑡 + 𝑒𝑡 t=1, …, n (13)

where 𝑌𝑡 is the observed data at time point t, 𝑇𝑡 is the trend component, 𝑆𝑡 is the
seasonal component, 𝑒𝑡 is the remainder, and 𝑇𝑡 is a piecewise linear function
with m+1 segments (Verbesselt & Hyndman 2010). Considering that each seg-
ment has a specific slope and intercept, there are m breakpoints 𝜏1∗ ,…, 𝜏𝑚

, such
that:

Tt = α̂ i + β̂ j t ∗
𝜏𝑖−1 < 𝑡 < 𝜏𝑖∗ (14)

̂i , and slope β̂j can be used to calculate the magni-


where 𝑖 = 1, … , m, intercept α
tude of the abrupt change and slope of the gradual change between detected
breaks points. Finally, the breakpoints of a trend at 𝜏1∗ ,…, 𝜏𝑚 ∗
can be estimated
using the residuals-based moving sum test and are assessed by minimizing a
Bayesian information criterion (BIC) from the seasonally adjusted data 𝑌𝑡 − 𝑆̂𝑡 ,
which is the iteration that represents the seasonal component using a nonpara-
metric season‐trend decomposition (STL) method (Cleveland et al. 1990). Sub-
sequently, the estimates of 𝑆̂𝑡 and 𝑌𝑡 will be calculated until the number and po-
sition of the detected breakpoints remain unchanged. After finding a breakpoint
in the time series, the mean of albedo values for 365 days before and 365 days
after the breakpoint was estimated (corresponding results can be seen in Fig. 9).

4.3.3. Albedo and forest structure

Although offering a comprehensive definition to characterize forest structure


might be quite challenging (Delang & Li 2012), in the context of this disserta-
tion, forest structure is represented by forest density, tree cover, and LAI, which
describe the distribution of forest structure and leaves/needles horizontally and
vertically. Forest density defines the number of trees per unit area. Tree cover is

37
defined as the proportional, vertically projected area of woody plants above a
given height (Sexton et al. 2013). LAI is defined in subchapter 2.1.2.
To explore the degree to which albedo can be explained with forest structure
on a global scale, the following three steps were taken: (1) daily albedo is calcu-
lated (subchapter 4.3.1) and other variables (forest structure and ancillary in-
formation) were obtained for the year 2005 (subchapters 3.2 and 3.3). Next (2),
I used a time series of LAI to estimate the timing of peak growing season at each
sub-biome using the Gu method (Gu et al. 2009), and I prepared a dataset of
high-quality albedo, forest structure, and ancillary information corresponding
to the timing of the peak growing season in 2005. Finally (3), I calculated the
links between forest structure and albedo using generalized additive models
(GAMs). Prior to developing the GAMs, it was tested whether the predictor var-
iables are subjected to multicollinearity, which showed that none of the predic-
tors were strongly correlated (Dormann et al. 2013).
The GAM, a flexible nonparametric approach (Hastie & Tibshirani 1990), was
fitted between albedo (response variable) and forest density, tree cover, and LAI
(predictor variables), as presented in Publication IV. The GAM provides
smoothing functions in dealing with complex and nonlinear patterns available
in a model (Hastie & Tibshirani 1990). The GAM was fitted (Wood 2017) as:

(15)

where 𝛼𝑏𝑠 stands for λ of shortwave, near-infrared and visible black sky albedo
is provided at 38° of SZA. The 𝛾0 parameter represents the model intercept,
while 𝑠1 , 𝑠2 , 𝑠3 represent smooth functions fitted using splines. The term ε is
equal to the residuals of the model, assuming that they follow a normal distri-
bution. To test whether the interaction between the forest structure variables
can also explain albedo, I also fitted new GAMs, to which the interactions be-
tween different pairs of the predictors were added as the new terms to the orig-
inal form of the model (see equation 3 of Publication IV). However, since the
interaction terms were not significant in the GAM models, I only used the mod-
els without the interaction terms (see Table 4 of Publication IV).
Next, a variance partitioning method was used to examine the proportion of
forest albedo explained by each forest structure variable. To avoid overfitting,
the penalized iteratively re-weighted least squares (P-IRLS) method (Wood
2000) was performed. Finally, I tested the spatially varying links between forest
structure and albedo by fitting a set of GAMs over a globally extended coarse
resolution (coarse grid cell: 50-km × 50-km) regular grid map. For that, each
coarse grid cell was considered as a spatial unit over which the GAM fitting was
carried out. Finally, the model’s goodness of fit (𝑅2 ) was assigned to the coarse
grid cell (corresponding results can be seen in Fig. 10).
I also explored the response of albedo to forest structure in each sub-biome.
For that, I quantified the variations of albedo in response to each forest struc-
ture variable followed by Theil-Sen’s slope method (Sen 1968), called Sen’s
slope hereafter (corresponding results can be seen in Table 5). Sen’s slope is a

38
Methods

non-parametric method and is used to estimate the slope of relationships be-


tween albedo and forest structure variables. Sen’s slope is considered as an effi-
cient method to treat the outliers and is calculated via (Sen 1968):

𝑦𝑖 −𝑦𝑗
𝑑𝑖𝑗 = 𝑚𝑒𝑑𝑖𝑎𝑛 (16)
𝑥𝑖 −𝑥𝑗

where 𝑦𝑖 and 𝑦𝑗 are data values at times 𝑥𝑖 and 𝑥𝑗 (or two consecutive values),
respectively with 𝑖 > 𝑗. The median of N values of 𝑑𝑖𝑗 is Sen’s estimator of the
slope. I then estimated the %change of the slope, followed by (Yue & Hashino
2003):

𝑚𝑒𝑑𝑖𝑎𝑛 𝑠𝑙𝑜𝑝𝑒×(𝑚𝑎𝑥𝑥 −𝑚𝑖𝑛𝑥 )


% 𝑐ℎ𝑎𝑛𝑔𝑒 = 𝑚𝑒𝑎𝑛
(17)

39
5. Results

In this chapter, the most significant results relevant to the objectives of this dis-
sertation (subchapter 1.5) are summarized. For more detailed results, interested
readers are encouraged to study the papers published on each particular topic.
In subchapter 5.1, the results of identifying ecosystem dynamics using temporal
analysis (Publication I) and the spatial and spatio-temporal analysis (Publica-
tion II) are presented. The results of the temporal dynamics of albedo in re-
sponse to disturbances (Publication III) are highlighted in subchapter 5.2, and
the studied global scale relationship between forest structure and albedo (Pub-
lication IV) is discussed in subchapter 5.3. Immediate comments on the results
are provided here and more detailed discussions can be found in Chapter 6.

Monitoring ecosystem dynamics

5.1.1. Temporal analysis of early warning signals (TEW)

The TEW approach (subchapter 4.2.1) may represent two types of alarms: a false
alarm that can either be a false negative (it is wrong to get no alarm) or a false
positive (it is wrong to get an alarm), and a true alarm that can either be a true
negative (it is correct to get no alarm) or a true positive (it is correct to get an
alarm). As is mentioned earlier, increases in the trend of autocorrelation, stand-
ard deviation, and skewness can be considered as an early warning signal of
critical transition in an ecosystem (Publication I).
For the Dorge Sangi wetland, analyzing the statistical signatures of autocorre-
lation, standard deviation, and skewness of NDVI yielded Kendall’s τ of up to
0.44, which was not as strong a positive trend as MVWR over time (Table 3).
The standard deviation of NDVI showed a negative trend by Kendall’s τ of -0.51
(false negative). The positive trend changes in statistical signatures of MNDWI
in unhealthy wetlands were also observed with Kendall’s τ of up to 0.40 (true
positive), except for the statistical signatures of standard deviation for which
Kendall’s τ of up to -0.78 was obtained (false negative) (Table 3). It implies that
some of the statistical signatures of NDVI and MNDWI can produce false nega-
tive alarms. I also investigated the performance of VWR in Publication I and
found that it also generated false negative alarms in the unhealthy wetland.
Among all studied state variables, the MVWR was the only one for which the

40
Results

trends in all the statistical signatures were significantly positive with Kendall’s
τ of up to 0.85 during the study period (Fig. 5, Table 3), resulting in a true pos-
itive alarm in the unhealthy Dorge Sangi wetland. These results may suggest the
MVWR as a potentially more robust state variable indicator to represent the
state of an unhealthy wetland.
The analysis results of the healthy study site, the Arpi wetland, showed no pos-
itive trend change in the statistical signatures of NDVI, MNDWI, and MVWR,
meaning that the majority of state variable indicators provided a true negative
alarm (Fig. 5, Table 3).

Figure 5. A time series of the statistical signatures of ecosystem state variables including NDVI,
MNDWI, and MVWR in the unhealthy Dorge Sangi (A) and healthy Arpi (B) wetlands. The red
dotted lines refer to the width of the rolling window. The blue dotted lines indicate the timing of
the trend changes. In each graph, the x-axis represents time. Figure adapted from Publication I.

41
Table 3. Kendall’s τ of trend changes in the statistical signatures (i.e., autocorrelation, standard
deviation, and skewness) of ecosystem state variables (i.e., NDVI, MNDWI, and MVWR) and its
associated significance (Dorge Sangi | Arpi). Table adapted from Publication I.

Indicators
Spectral indices
Autocorrelation Standard deviation Skewness

NDVI 0.30* | -0.29*** -0.51***|-0.40*** 0.44* | -0.41***


MNDWI 0.40***| -0.27*** -0.78***|-0.50*** 0.13***| 0.27

MVWR 0.85***| -0.22*** 0.41***| 0.28*** 0.75***| -0.02***


Significance level: *** P < 0.001, ** P < 0.01, * P < 0.05, and P < 0.1

5.1.2. SPT, STG, and STM analysis

The magnitude of changes in NDVI showed that the unhealthy Yosemite and
Sequoia forest sites experienced NDVI decreases in some forest patches for up
to -0.6 units, which was higher than that of the NDVI in the healthy Wood Buf-
falo forest site (Fig. 6). The patches that experienced a decrease in NDVI (Fig.
6) also showed an increase in the results of STG and decrease in STP (Fig. 7).
This infers to local spatial autocorrelation reductions and signals of reduced re-
silience, meaning both methods provided true positive alarms in unhealthy
study sites. The opposite scenario is true for the evidence ecosystem, where the
STM and STG provided true negative alarms. Furthermore, the results of the
SPT method showed a false positive alarm in the healthy Wood Buffalo study
site, meaning that those pixels that did not experience NDVI changes (Fig. 6)
showed early warning signals of state change.

42
Results

Yosemite Sequoia Wood Buffalo

Figure 6. The magnitude of the trend in the normalized difference vegetation index (NDVI) time
series between 2000 and 2019, calculated using Kendall’s τ. Kendall’s τ of zero means no signif-
icant trend in data, while a value of +0.6 or -0.6 (green color and red color in the figures) shows
an increasing or decreasing trend in NDVI, respectively. In each graph, the x-axis represents
longitude and the y-axis represents latitude. Figure adapted from Publication II.

43
Yosemite Sequoia Wood Buffalo

Figure 7. Early warning signals of forest mortality in the Yosemite, Sequoia, and Wood Buffalo
national parks. The pixel values represent Kendall’s τ of the statistical signature of normalized
vegetation index (NDVI). Kendall’s τ of zero means no significant trend in the corresponding data,
while a value of +1 or -1 (red color and dark green color in the figures) shows an increasing or
decreasing trend in the corresponding data, respectively. The first row represents STM analysis,
the second row represents the results of the STG method, and the third row represents the spatial
pattern of temporal analysis of early warning signals (lag-1 autocorrelation), i.e., SPT analysis. In
each map, the x-axis represents longitude and the y-axis represents latitude. Figure adapted from
Publication II.

The temporal dynamics of forest albedo

The results of large-scale studies (the entire Zagros forest) showed a strong in-
verse correlation between the albedo and the number of fire events and LAI, as
well as a weak correlation between the mean monthly precipitation and burn
severity for the period between 2000 to 2016 (Fig. 8, Table 4).

44
Results

For the BFAST method results, the largest significant breakpoints in the al-
bedo time series were observed in 2007 and 2008 in the Dashte baram and Ma-
rivan study sites, respectively. This seems to be due to the occurrence of ENSO
in the Middle East, which caused a severe drought in the Zagros area between
2007 and 2008 and which abruptly affected the ecosystems in the region (Trigo
et al. 2010; Voss et al. 2013; Hosseinzadeh Talaee et al. 2014; Barlow et al.
2016). On the other hand, in Ilam and Ize, the largest breakpoints in the tem-
poral dynamics of albedo occurred in 2013 and 2014, where severe fires were
the main forest disturbance (Fig. 9).
Regarding the magnitude of albedo changes in the four study sites, the results
showed that whenever the fire events were the primary disturbance in Ilam and
Ize, it resulted in a reduction in the mean albedo. However, whenever the
drought was the primary disturbance driver (e.g., in Marivan and Dashte
baram) an increase in the mean albedo became noticeable (Fig. 9).

Figure 8. A–B: The time series of black sky albedo (grey color) and leaf area index (green
color) in right and left y-axes, respectively; C–D: The time series of mean monthly precipitation
(blue color) and the number of fire events per month (brown color) in left and right y-axes, respec-
tively. The x-axes refer to time. Figure adapted from Publication III.

45
Figure 9. Breakpoints in the time series of black sky albedo in the four study sites based on
the break detection for additive season and trend (BFAST) method. In each graph, the black line
represents the fitted model and the blue line represents the trend. The actual data and breakpoint
occurrence points are specified with the gray and vertical dashed lines, respectively. The x-axes
refer to time. Figure adapted from Publication III.

46
Results

Table 4. The links between albedo and mean monthly precipitation, the number of fire events per
month, burn severity, and leaf area index (LAI), aggregated (using the mean function) in the
northern and southern Zagros between 2000 and 2016. Table adapted from Publication III.

Study region Correlation coefficient

number of
precipitation burn severity
fire events & LAI & albedo
& albedo & albedo
albedo

Northern Zagros -0.27* -0.40* 0.18* -0.74*

Southern Zagros -0.25* -0.59* 0.17 -0.69*

The asterisk (*) represents the significant correlation test at 0.01 using the p-value. The levels
of burn severity are classified as low severity if it ranges between +0.1 to +0.27, moderate severity
between +0.27 to +0.66, and high severity at greater than +0.66.

Albedo and forest structure

The findings confirmed the varying relationships between forest structure and
albedo pixelwise across the globe (Fig. 10). The relationships between forest
structure and visible albedo showed an average 𝑅2 of 0.40, which was only
slightly weaker than shortwave and near-infrared albedos with average 𝑅2 of
0.43 (Fig. 10). The 𝑅2 values in the map were higher in the northern regions of
the boreal zone (e.g., North America and central Eurasia) compared with other
areas (Fig. 10).
While increases in near-infrared and shortwave albedos with increasing leaf
area index were apparent in all the forests, the relationships between tree cover
and near-infrared and shortwave albedos were opposite for all the studied for-
ests, except for the DB forests in the mediterranean and tropical regions (Table
5). The % change of albedo in response to forest density showed diverse varia-
tion in different forests around the world (Table 5). According to the results,
visible albedo decreased globally in response to increasing forest structure var-
iables in the majority of the forests.

47
A

Figure 10. The spatial distribution of the 𝐑𝟐 of the generalized additive models (GAMs) repre-
senting the links between forest structure and (A) shortwave, (B) near-infrared, and (C) visible
albedos during the peak growing season in 2005. The grey shades represent the mean 𝑹𝟐 values
over longitudes (top) and latitudes (right), respectively. Figure adapted from Publication IV.

48
Results

49
Table 1. The % change of albedos in response to forest structure variables (i.e., forest density, tree cover, and leaf area index) changes. The columns refer to forest types,
that include evergreen needleleaf (EN), evergreen broadleaf (EB), deciduous needleleaf (DN), deciduous broadleaf (DB), mixed forests (Mixed), and woody savannah forests
(WS). All the trends have been significant at the level of 90%. Table adapted from Publication IV.
shortwave albedo
density tree cover LAI
EN EB DN DB Mix WS EN EB DN DB Mix WS EN EB DN DB Mix WS
boreal 17 NA 19 -18 16 9.3 -34 NA -24 -19 -24 -6 57 NA 2 50 41 55
mediterranean -13 16 NA -4 8 4.7 -20 -15 NA 4 -9 -25 12 50 NA 6.2 57 74
temperate 8.1 -8 -9 11.3 6 -10 -26 -29 -6 5 -19 -13 41 31 10 3 25 50
tropical -22 -20 NA -29 -16 -14 -12 -14 NA -3 -19 -6 59 10 NA 39 29 27
near-infrared albedo
EN EB DN DB Mix WS EN EB DN DB Mix WS EN EB DN DB Mix WS
boreal 24 NA -2 -3 -13 34 -43 NA -25 -15 -35 -43 65 NA 4 78 54 67
mediterranean -14 34 NA -5 13 14 -25 -11 NA 9 0 -24 31 45 NA -7 62 94
temperate -2 30 6 2 -6 -9 -25 -33 -2 7 -10 -21 52 44 14 3 30 67
tropical -25 -20 NA -41 3 -32 0 -15 NA -1 -24 -7 70 29 NA 59 57 46
visible albedo
EN EB DN DB Mix WS EN EB DN DB Mix WS EN EB DN DB Mix WS
boreal -17 NA 0 0 0 12 -64 NA -20 -32 -33 -46 -8 NA 1 -48 -13 -26
mediterranean -26 2 NA -3 -5 -9 -21 -82 NA -57 -50 -40 -77 37.7 NA -17 -6 -46
temperate 0 0 -10 -3 7 11 -47 -72 -17 -32 -40 -24 1 -46 2.1 -4 -9 -21
tropical -65 -28 NA -49 5 -32 -99 -44 NA 41 -4 -11 -11 -18 NA -105 -73 -58
6. Discussion

Monitoring ecosystem dynamics

6.1.1. Ecosystem state variables

In Publication I, in line with Objective I, the chief focus was on exploring the
early warning signals of a critical transition in both a healthy and an unhealthy
wetland. It was shown that, by using an appropriate ecosystem state variable
obtained from satellite images, reliable early warning signals of a critical tran-
sition could be detected, and the risk of false alarms can be decreased. To this
end, a new spectral composite, namely MVWR, was developed. MVWR assists
in generating robust early warning signals of state changes in wetlands.
Applying appropriate ecosystem state variables can play a significant role in
generating accurate early warning signals, as well as reducing the risk of false
alarms by the indicators. As was discussed in Publication I, using NDVI solely
to monitor the state of an unhealthy wetland generated a false negative alarm
while a true positive alarm is expected (Fig. 5, Table 3).
The MVWR was shown to be a robust ecosystem state variable in monitoring
the state of wetlands because it showed true positive alarms regarding state
changes in unhealthy wetlands with Kendall’s τ of up to 0.85 and provided true
negative alarms for the healthy wetland with Kendall’s τ of up to 0.28. MVWR
is estimated using four spectral bands, including red, near-infrared, green, and
short infrared spectral bands. Red and near-infrared bands can be used to infer
the total amount of vegetation. Green and SWIR bands, on the other hand, can
be used to infer the water area. These four bands contain more information
about the wetland than the single vegetation-based indicator or water-based in-
dicator, which may explain the reason behind the potentially better perfor-
mance of MVWR to characterize the dynamics of a wetland, compared to NDVI
and MNDWI (Zhao et al. 2009).

6.1.2. Methods of identifying ecosystem dynamics

To benchmark the new approach, two unhealthy forest ecosystems with a


known history of resilience decline (Liu et al. 2019) were studied. For both eco-
systems, extensive field plots confirmed sudden forest mortalities between 2010
50
Discussion

and 2013 (Liu et al. 2019), which was matched with the breakpoint occurrence
in the time series of NDVI using the BFAST method. It is shown that the STM
and STG approaches outperform TEW and SPT in detecting forest mortality two
years prior to its occurrence.
According to the results, applying the TEW approach yielded false negative
alarms in unhealthy forests between 2000 and two years prior to the occurrence
of abrupt forest mortality (Fig. 5 in Publication II). One main reason for these
false alarms is attributed to the distribution of different forest types with varying
resilience across the study areas (Liu et al. 2019). Note that, by aggregating the
signals of different forest types in the TEW approach, the negative and positive
values of signals may cancel each other out, resulting in false alarms. To over-
come the problem of aggregation in the TEW approach, I performed SPT anal-
yses at study sites. The SPT analyses showed some true positive alarms of forest
mortality in unhealthy study sites. However, it also showed false positive alarms
in the healthy forest site (see the comparison between Fig. 6 and Fig. 7). The
possible reason could be related to the estimation of statistical signatures in the
TEW and SPT approaches that are shown to be affected by the choice of window
size and preprocessing. Furthermore, the results of the TEW and SPT ap-
proaches can be influenced by detrending methods (TAMGA 2019). Another
possible reason could be related to some unhealthy forests that did not experi-
ence abrupt change between 2010 and 2013, and this is why I obtained early
warning signals of possible upcoming critical transitions for these pixels.
The new methods (STM and STG) showed their efficiency to overcome the
abovementioned problems. According to the results, the newly proposed ap-
proach could reliably identify early warning signals two years prior to abrupt
forest mortality (see the comparison between Fig. 6 and Fig. 7). The possible
reason behind the robustness of the STM and STG methods in generating early
warning signals of forest mortality is that these methods do not require pre-
processing steps that are essential with the TEW and SPT approaches (e.g., us-
ing a rolling window with a selection of arbitrary size). The robustness of gener-
ating early warning signals using spatial indicators such as Moran’s 𝐼 has been
previously reported (Dakos et al. 2015).

The temporal dynamics of albedo

In Publication III, Objective II of this dissertation was addressed. For that, the
degree to which disturbances can explain albedo dynamics of a sparse forest
area in the Middle East was explored. It was shown that albedo dynamics are
strongly linked to the number of fire events and LAI (Fig. 8, Table 4). Also, pre-
cipitation rate and burn severity were found to only weakly explain forest albedo
(Fig. 8, Table 4). Furthermore, I found that fire and drought (a lack of precipi-
tation) can, respectively, abruptly decrease and increase the albedo (Fig. 9).
For the rate of precipitation change, the results showed that there is a rela-
tively weak and negative link between albedo and precipitation with a correla-
tion coefficient of -0.27 and -0.25 in the northern and southern Zagros forests,
respectively (Fig. 8, Table 4). According to the results, the largest albedo

51
changes in drought-affected forests (Marivan and Dashte baram) happened for
El Niño/Southern Oscillation (ENSO) governed conditions in the whole Middle
East, resulting in a severe drought (Fig. 9). The abrupt changes in ecosystems
due to ENSO have also been reported in several other studies (Trigo et al. 2010;
Voss et al. 2013; Hosseinzadeh Talaee et al. 2014; Barlow et al. 2016). According
to the results of Publication III, the response of the forests in the Marivan and
Dashte baram sites to severe drought conditions was abrupt increases in forest
albedo by 0.002 and 0.004, respectively (Fig. 9). Increasing albedo values can
happen due to defoliation and canopy cover losses by drought, resulting in more
exposure of the spectral signature of soils that can be optically brighter than
overstory (Govaerts & Lattanzlo 2007).
Concerning the relationships between the number of fire events and albedo,
the results showed a relatively strong connection between these two parameters,
with a correlation coefficient of -0.40 and -0.59 in the northern and southern
Zagros, respectively (Fig. 8, Table 4). Furthermore, for the Ilam and Ize sites,
for which fire was the main disturbance, the most significant abrupt albedo de-
crease was −0.013, and −0.028, respectively (Fig. 9). A similar decreasing trend
of albedo in forests upon a fire regime by -0.02 has also been reported (Jin &
Roy 2005). The decreasing trend in albedo following a fire regime can be at-
tributed to the optically dark black carbon released after a fire event that stays
on the surface. Dark black carbon can enhance the absorption of incoming ra-
diation and, hence, reduce the albedo. Dark black carbon produced after a fire
event typically remains until it is either washed out by precipitation or is covered
by plant regrowth. The albedo reduction by forest fire has already been reported
in a number of studies (Veraverbeke 2010; Dintwe et al. 2017).
The results of Publication III also showed that burn severity could explain al-
bedo with a correlation of 0.18 and 0.17 in the northern and southern Zagros,
respectively (Fig. 8, Table 4). The reason for such a low correlation is the high
diversity of the response of understory and overstory species to disturbances
(Fattahi & Ildoromi 2011; Pourreza et al. 2014; Heydari et al. 2017). In other
words, this might be related to the varying feedback of the Zagros forest to the
different degrees of burn severity, which makes the predictability of albedo in
response to burn severity more challenging (Pourreza et al. 2014).
I also observed a strong negative relationship between albedo and LAI with a
correlation coefficient of -0.74 in the northern Zagros and -0.69 in the southern
Zagros (Fig. 8). I discuss the relationships between albedo and LAI further in
the next subchapter.

Albedo modulation by forest structure

In Publication IV, I focused on the role forest structure has in modulating forest
albedo, which was Objective III of this dissertation. For that, I explored the re-
lationship between albedo and forest structure. I showed that forest structure
could modulate albedo in the majority of the forests around the world. Publica-

52
Discussion

tion IV was carried out spatially explicit, as previous studies highlighted the im-
portance of geographic locations in albedo studies (Hovi et al. 2019, 2016; Lukeš
et al. 2014).
In this study, the overall changes in the 𝑅2 values of the relationship between
forest structure and albedo during the peak growing season in 2005 across lati-
tudes and longitudes were mapped (grey shades surrounding the images in Fig.
10). As can be seen in Fig. 10, the mean 𝑅2 values over longitudes and latitudes
showed no significant pattern as also previously reported by Lukeš et al. (2014).
The weak relationships across latitudes and longitudes occurred due to having
both high and low values of 𝑅2 in forests across the latitude and longitude val-
ues, which may cancel each other out by aggregating using the mean function.
The pixelwise spatial distribution of 𝑅2 values showed substantially varying
2
𝑅 values over different geographic locations (Fig. 10). For example, the spatial
maps indicated that the 𝑅2 values of the relationships between forest structure
and albedo were higher in the northern parts of the boreal region compared with
the other regions (Fig. 10). Varying values of 𝑅2 pixelwise may explain the in-
consistent results reported for the links between forest structure in the literature
(Sun et al. 2010; Dore et al. 2012; Lukeš et al. 2013a). Such varying 𝑅2 values
seem to be related to the varying values of forest structure and having different
forest types. It can also be influenced by the varying proportions of forest floor,
including different species composition and visible understory in each pixel,
which can simultaneously affect albedo. This suggests that the relationships be-
tween forest structure and albedo are very much dependent on the geographical
location, as also shown by Hovi et al. (2019) for the boreal region. Furthermore,
the findings of Publication IV emphasize the importance of spatially explicit per-
formance of the analysis of the links between land surface albedo and forest
structure.
Another outcome of this study implies that for the forests for which the corre-
lation between forest structure and albedo yield high 𝑅2 values, e.g., evergreen
needleleaf forests (Table 4 of Publication IV and Fig. 10), albedo should be con-
sidered when making forest management plans. Accordingly, for other regions
where forest structure and albedo are not so tightly connected, albedo can be
overlooked in forest management plans.
In this work, the magnitude of response of albedo to each forest structure var-
iable in each sub-biome was also studied. The results showed a varying %change
in forests, even within the same forest types (Table 5). This seems to be due to a
varying range of forest structure variables, or the confounding effects of optical
properties of overstory and forest floor components. More specifically, the
%change in albedo as a function of forest density showed distinctly different
values in various sub-biomes. This implies that forest density alone cannot ex-
plain forest albedo on a global scale, although it is one of the variables used in
the physical modeling of albedo (e.g., Wang 2005). The results also indicated
negative relationships between the tree cover and albedo in forests around the
world (except for DB in the mediterranean and temperate regions, as reported
in Table 5). This can be attributed to the lower reflectance of tree cover com-
pared with bare soil (Bonan 2008; Zhao & Jackson 2014), which consequently

53
can decrease albedo due to higher absorption (Planque et al. 2017). The results
of Publication IV also showed a positive relationship between albedo and LAI,
which is opposite to the results of Publication III, where I observed a robust
negative relationship during a year. The possible reason for this observation is
that in some forest environments, developing tree-layer vegetation can decrease
surface albedo via eliminating the effects of forest floor which itself might have
a higher albedo than overstory vegetation (Tian et al. 2018).

Uncertainties and recommendations for future studies

Detailed information about the sources of limitations and uncertainties of each


study has been discussed in each publication. In this subchapter, the main lim-
itations, as well as some general recommendations for their handling in future
studies, are summarized.
One of the main challenges in the conducted studies throughout this disserta-
tion was the cloud impact on the acquired data, which could be a hindrance to
fully achieving the chief objectives. The fusion of optical satellite images with a
new generation of radar mission sentinel data (sentinel 1 and 3) may enable us
to achieve cloud-free data in the future and overcome the limitations of optical
satellite images. Until that moment, careful preprocessing of data is highly rec-
ommended to produce reliable results in albedo studies. Preprocessing of al-
bedo is especially crucial for forest albedo analysis. In general, land surface al-
bedo, depending on the materials, can have substantially varying values. It can
be zero for blackbody, which is an idealized physical body that absorbs all inci-
dent radiation, to around 0.9 for fresh and very small-grained pure snow
(Riihelä 2013). Nevertheless, albedo variations in forests have a reasonably nar-
row range that is up to the value of 0.13 in the majority of forests around the
world during peak growing season (Table A5 of Publication IV). In some sparse
forests, the albedo variation is even narrower than other forests, i.e., around al-
bedo of 0.02 (Table 2 of Publication III). Therefore, very low variations in forest
albedo can imply the importance of taking special care with preprocessing steps,
e.g., keeping only high-quality pixels and taking care of slope and water influ-
ences on albedo values, as well as the uncertainties of albedo products. For ex-
ample, a new albedo satellite product, namely LSA-SAF/EUMETSAT derived
from the EPS/METOP AVHRR sensor2, has just recently been released
(2019/05/06). In this product, relative uncertainty for albedos smaller than
0.15 is 0.015 in absolute albedo units. This means that forest albedo, which is
often smaller than 0.15 during peak growing season, can be subjected to high
uncertainty.
Another challenge that can be considered as a source of uncertainty in this
dissertation is the lack of ground truth data to validate the findings. Considering
the current political situation in some study areas (e.g., the Middle East sites),

2
https://landsaf.ipma.pt/en/products/albedo/etal, accessed data: 25.05.2020

54
Discussion

accessing field data is almost infeasible. As a result, in Publications I and III,


satellite data was almost the only way to perform large scale analysis. For that,
very high-resolution satellite images (e.g., IKONOS, Pléiades, and Quickbird)
were used to support the findings in Publication III, which can be recommended
as an alternative solution for such study sites in future studies.
Another source of limitations in this study was related to using the satellite
images with medium spatial resolution in the analyses (e.g., usage of 500-m and
1-km spatial resolution in Publications I–IV). The concept of scale is a key issue
in the different fields of studies (Liang 2000). For example, the scale depend-
encies of the results of albedo measurements have been studied by Kuusinen
(2014). The scale dependencies of the results restrict the validity domain of the
related findings to only the spatial resolution considered in the studies. It lim-
ited the inferences drawn from the results of the studies performed in this dis-
sertation to another scale. Another issue related to the spatial resolution is re-
lated to its importance for identifying ecosystem dynamics. It has been reported
that the rate of success in detecting reliable early warning signals is considerably
dependent on the spatial resolution of the data (Nijp et al. 2019). For example,
if the images are of coarse spatial resolution, the signals of small-scale changes
may vanish or can be canceled out. So far, the efficiency of the 30-m spatial res-
olution (using Landsat images by Liu et al. (2019)) to 500-m spatial resolution
(using MODIS images in Publications I and II) in identifying ecosystem dynam-
ics have been demonstrated. As a future perspective, multispectral bands of new
generations of satellite images, such as Sentinel-2A with 13 spectral bands
(443–2190 nm) and 10m to 60m spatial resolution, may open up opportunities
to assist with monitoring ecosystem dynamics by providing global high spatio-
temporal resolution data (Sentinel 2013). Monitoring ecosystem dynamics usu-
ally requires long term data, however, and Sentinel-2, operating since June 23,
2015, still has not provided sufficient observations for identifying ecosystem dy-
namics. Accessing the adequate archive of these satellite images, which is ex-
pected in the few next years, can considerably facilitate future studies in the
real-time spatio-temporal monitoring of ecosystem dynamics.
Another limitation of the results of this study is related to the limited number
of study sites in Publication I and Publication II. For example, Publication I in-
cluded limiting the study only to the wetland that was disturbed with drought
conditions. Publication II only focused on three drought-affected evergreen
needleleaf forests in national parks. Obviously, for a broader application, eval-
uating the performance of the methods on different forests with different forest
types and climate conditions would be more advantageous. As another interest-
ing topic for future studies, the fusion of ecosystem change detection methods
(see the review by Zhu (2017)), with the methods to explore ecosystem dynam-
ics, is recommended. For example, a robust change detection method, such as
BFAST, can be performed on a global scale to detect the pixels that experienced
abrupt changes in intact ecosystems. Then, ecosystem dynamics approaches can
be examined to measure whether the early warning signals of ecosystem col-
lapse could truly be captured in a timely manner.

55
The lessons from Publication II and Publication IV tell us that further at-
tempts to monitor the variation of a target variable (e.g., albedo) and ecosystem
dynamics using satellite images can be improved by spatially explicit analysis.
As discussed before, aggregating all the data from the entire study site may pro-
vide misleading information and vanish the variation signals in an ecosystem or
albedo variations. There are already some developed tools and methods for spa-
tially explicit monitoring of ecosystems, such as bfastSpatial3 (Verbesselt et al.
2012b) that can facilitate supporting pixelwise analysis.

The fusion of knowledge about ecosystem state dynam-


ics theory with albedo dynamics

Although throughout this dissertation I discussed two distinct concepts, namely


ecosystem dynamics and albedo dynamics, in reality, they are inherently linked.
An example of this is the different feedbacks of albedo to disturbances in the
southern and northern parts of the Zagros forest due to the diversity of resili-
ence of the ecosystems in those study sites (Publication III).
More specifically, previous studies showed that there is a strong link between
the severity of disturbances (e.g., drought and fungi and beetle attacks) with
forest defoliation and forest loss (Capretti & Battisti 2007; Vannini et al. 2009;
Linaldeddu et al. 2011). In Publication III, it was shown that LAI, which repre-
sents the effects of the severity of disturbances on forest defoliation and forest
loss, is an important and explanatory factor of albedo in study sites with a cor-
relation coefficient of -0.69 and -0.74 (Table 4). For example, where the de-
crease in LAI was greater (Table 5 of Publication III), the magnitude of changes
in albedo was also greater. It suggests that we can determine the intensity of
albedo variations in forests by monitoring the rate of vegetation loss upon dis-
turbances that can itself be controlled with the resilience of ecosystems, noting
that vegetation-based indicators such as NDVI/LAI are shown to be appropriate
ecosystem state variables in Publication II. These all suggest the further study
of land surface albedo as a potentially more efficient ecosystem state variable in
future studies.

3
https://www.rdocumentation.org/packages/bfastSpatial/versions/0.6.2, accessed date:
27.05.2020

56
Summary

7. Summary

This dissertation explored the advantages of using optical satellite remote sens-
ing to understand the dynamics of an ecosystem (forests and wetlands were
used as the case studies) to provide insight into the state of an ecosystem and its
response to disturbances. Half of the dissertation was dedicated to enhancing
knowledge about the changes in land surface albedo, a key ecosystem variable
that can be used to study the dynamics and state of an ecosystem (Fig. 11).
Publications I and II explore how temporal and spatio-temporal remotely
sensed data can improve our understanding of ecosystem dynamics. It was
shown (Publication I) that using an appropriate state variable in a wetland eco-
system, which is linked to the key processes of the ecosystem, enhance our abil-
ity to measure the resilience of the ecosystem and generate early warning signals
of critical transitions. A new remotely sensed spectral composite index was pro-
posed (MVWR) as the key ecosystem variable that represents the dynamics of
both vegetation and water contents in wetlands. Publication II proposed a new
approach to understanding the risk of reduced resilience in forests by exploring
the dynamics of changes in local spatial structure derived from spatio-temporal
satellite images. These two publications improved our ability to generate early
warning signals of an ecosystem state change and also provided a convenient
implementation of the presented approach as well as other state-of-the-art met-
rics in a new R package, “stew”.
Further, the response of forest ecosystems to disturbances was investigated in
Publications III and IV, where land surface albedo was specifically considered
as a key variable linked to the ecosystem properties (e.g., forest structure). It
was shown that disturbances (the rate of precipitation, linked to droughts, and
the number of fire events) alter the dynamics of land surface albedo (Publication
III). In addition, the links between forest structure and albedo were investigated
to understand the role of forests in modulating land surface albedo across dif-
ferent biomes on a global scale (Publication IV). This study then provided a
global dataset representing the links between albedo and forest structure and
more information about the effects of solar zenith angle on albedo in a separate
publication by Alibakhshi et al. (2020), as well as a new web-application that
enables users to go through the results of Publication IV: https://al-
bedo.shinyapps.io/shiny_apps/.

57
Figure 11. A conceptual model that represents the impacts of climate change-induced disturb-
ances on ecosystem components. The dotted square represents the ecosystem dynam-
ics. The grey shapes and their links, together with the dotted square, are partly investi-
gated in Publications I to IV. The so-called “other parameters” refer to the other key bio-
geophysical parameters (e.g., evapotranspiration, surface roughness), biochemical com-
ponents (e.g., carbon cycle), soil moisture, etc. This model shows that global and local
climate change can invoke climate-induced disturbance effects (Publications I to III). Dis-
turbances, in turn, can affect ecosystem properties that can be represented by key varia-
bles, such as land surface albedo (Publication III) and forest structure (Publication IV).
These changes can lead to ecosystem resilience loss that can provoke a critical transition
(Publications I and II).

58
References

8. References

Adams, H.D., MacAlady, A.K., Breshears, D.D., Allen, C.D., Stephenson, N.L.,
Saleska, S.R., et al. (2010). Climate-induced tree mortality: Earth system
consequences. Eos (Washington. DC)., 91, 153–154.
Ahmadi, R., Mohebbi, F., Hagigi, P., Esmailly, L. & Salmanzadeh, R. (2011).
Macro-invertebrates in the Wetlands of the Zarrineh estuary at the south of
Urmia Lake (Iran). Int. J. Environ. Res., 5, 1047–1052.
Alibakhshi, S., Crowther, T.W. & Naimi, B. (2020a). Land surface black-sky
albedo at a fixed solar zenith angle and their relations to forest structure
during peak growing season based on remote sensing data. Data Br., 105720.
Alkama, R. & Cescatti, A. (2016). Biophysical climate impacts of recent
changes in global forest cover. Science (80-. )., 351, 600–604.
Allen, C.D., Macalady, A.K., Chenchouni, H., Bachelet, D., McDowell, N.,
Vennetier, M., et al. (2010). A global overview of drought and heat-induced
tree mortality reveals emerging climate change risks for forests. For. Ecol.
Manage., 259, 660–684.
Angeler, D.G. & Allen, C.R. (2016). Quantifying resilience. J. Appl. Ecol., 53,
617–624.
Angelsen, A. (2009). Realising REDD+: National strategy and policy
options. Cifor. ISBN: 978-6-02-869303-5.
Anselin, L. (1995). Local indicators of spatial association—LISA. Geogr.
Anal., 27, 93–115.
Barlow, M., Zaitchik, B., Paz, S., Black, E., Evans, J. & Hoell, A. (2016). A
review of drought in the Middle East and southwest Asia. J. Clim., 29, 8547–
8574.
Barnosky, A.D., Matzke, N., Tomiya, S., Wogan, G.O.U., Swartz, B., Quental,
T.B., et al. (2011). Has the Earth’s sixth mass extinction already arrived?
Nature, 471, 51–57.
Betts, A.K. & Ball, J.H. (1997). Albedo over the boreal forest. J. Geophys. Res.
Atmos., 102, 28901–28909.
Betts, R.A. (2000). Offset of the potential carbon sink from boreal forestation
by decreases in surface albedo. Nature, 408, 187.
Bonan, G.B. (2008). Forests and climate change: Forcings, feedbacks, and
the climate benefits of forests. Science (80-. )., 320, 1444–1449.
Bright, R.M., Antón‐Fernández, C., Astrup, R., Cherubini, F., Kvalevåg, M. &

59
Strømman, A.H. (2014). Climate change implications of shifting forest
management strategy in a boreal forest ecosystem of Norway. Glob. Chang.
Biol., 20, 607–621.
Bright, R.M., Davin, E., O’Halloran, T., Pongratz, J., Zhao, K. & Cescatti, A.
(2017). Local temperature response to land cover and management change
driven by non-radiative processes. Nat. Clim. Chang., 7, 296.
Bright, R.M., Eisner, S., Lund, M.T., Majasalmi, T., Myhre, G. & Astrup, R.
(2018). Inferring surface albedo prediction error linked to forest structure at
high latitudes. J. Geophys. Res. Atmos., 123, 4910–4925.
Capretti, P. & Battisti, A. (2007). Water stress and insect defoliation promote
the colonization of Quercus cerris by the fungus Biscogniauxia mediterranea.
For. Pathol., 37, 129–135.
Carpenter, S.R. & Brock, W.A. (2011). Early warnings of unknown nonlinear
shifts: A nonparametric approach. Ecology, 92, 2196–2201.
Carpenter, S.R., Cole, J.J., Pace, M.L., Batt, R., Brock, W.A., Cline, T., et al.
(2011). Early warnings of regime shifts: A whole-ecosystem experiment.
Science (80-. )., 332, 1079–1082.
Ceccato, P., Flasse, S., Tarantola, S., Jacquemoud, S. & Grégoire, J.-M.
(2001). Detecting vegetation leaf water content using reflectance in the
optical domain. Remote Sens. Environ., 77, 22–33.
Cescatti, A., Marcolla, B., Vannan, S.K.S., Pan, J.Y., Román, M.O., Yang, X.,
et al. (2012). Intercomparison of MODIS albedo retrievals and in situ
measurements across the global FLUXNET network. Remote Sens. Environ.,
121, 323–334.
Chambers, S.D. & Chapin, F.S. (2002). Fire effects on surface-atmosphere
energy exchange in Alaskan black spruce ecosystems: Implications for
feedbacks to regional climate. J. Geophys. Res., 108, 8145.
Chance, C.M., Hermosilla, T., Coops, N.C., Wulder, M.A. & White, J.C.
(2016). Effect of topographic correction on forest change detection using
spectral trend analysis of Landsat pixel-based composites. Int. J. Appl. Earth
Obs. Geoinf., 44, 186–194.
Charney, J.G. (1975). Dynamics of deserts and drought in the Sahel. Q. J. R.
Meteorol. Soc., 101, 193–202.
Chen, J.M. & Black, T.A. (1991). Measuring leaf area index of plant canopies
with branch architecture. Agric. For. Meteorol., 57, 1–12.
Chen, X., Vierling, L., Deering, D. & Conley, A. (2005). Monitoring boreal
forest leaf area index across a Siberian burn chronosequence: a MODIS
validation study. Int. J. Remote Sens., 26, 5433–5451.
Clara, P.E. da. (2013). Decline of mediterranean oak trees and its association
with Phytophtora cinnamomi: a review. Eur. J. For. Res., 132, 411–432.
Cleveland, R.B., Cleveland, W.S., McRae, J.E. & Terpenning, I. (1990). STL:
A seasonal-trend decomposition procedure based on loess. J. Off. Stat., 6, 3–
73.
Crowther, T.W., Glick, H.B., Covey, K.R., Bettigole, C., Maynard, D.S.,
Thomas, S.M., et al. (2015). Mapping tree density at a global scale. Nature,
525, 201.

60
References

Dakos, V., Carpenter, S.R., Brock, W.A., Ellison, A.M., Guttal, V., Ives, A.R.,
et al. (2012a). Methods for detecting early warnings of critical transitions in
time series illustrated using simulated ecological data. PLoS One, 7, e41010.
Dakos, V., Carpenter, S.R., Brock, W.A., Ellison, A.M., Guttal, V., Ives, A.R.,
et al. (2012b). Methods for detecting early warnings of critical transitions in
time series illustrated using simulated ecological data. PLoS One, 7, e41010.
Dakos, V., Carpenter, S.R., van Nes, E.H. & Scheffer, M.M. (2015). Resilience
indicators: Prospects and limitations for early warnings of regime shifts.
Philos. Trans. R. Soc. B Biol. Sci., 370, 1–10.
Dakos, V., Scheffer, M., Van Nes, E.H., Brovkin, V., Petoukhov, V. & Held, H.
(2008). Slowing down as an early warning signal for abrupt climate change.
Proc. Natl. Acad. Sci., 105, 14308–14312.
Delang, C.O. & Li, W.M. (2012). Ecological succession on fallowed shifting
cultivation fields: A review of the literature. Springer Science & Business
Media. ISBN: 9789400758216
Delbart, N., Kergoat, L., Le Toan, T., Lhermitte, J. & Picard, G. (2005).
Determination of phenological dates in boreal regions using normalized
difference water index. Remote Sens. Environ., 97, 26–38.
DeVries, B., Verbesselt, J., Kooistra, L. & Herold, M. (2015). Robust
monitoring of small-scale forest disturbances in a tropical montane forest
using Landsat time series. Remote Sens. Environ., 161, 107–121.
Didan, K. (2015). MOD13A1 MODIS/Terra Vegetation Indices 16-Day L3
Global 500m SIN Grid V006 [Data set], NASA EOSDIS LP DAAC.
Dintwe, K., Okin, G.S. & Xue, Y. (2017). Fire-induced albedo change and
surface radiative forcing in sub-Saharan Africa savanna ecosystems:
Implications for the energy balance. J. Geophys. Res. Atmos., 122, 6186–
6201.
Dore, S., Montes‐Helu, M., Hart, S.C., Hungate, B.A., Koch, G.W., Moon,
J.B., et al. (2012). Recovery of ponderosa pine ecosystem carbon and water
fluxes from thinning and stand‐replacing fire. Glob. Chang. Biol., 18, 3171–
3185.
Dormann, C.F., Elith, J., Bacher, S., Buchmann, C., Carl, G., Carré, G., et al.
(2013). Collinearity: a review of methods to deal with it and a simulation
study evaluating their performance. Ecography (Cop.)., 36, 27–46.
Drake, J.M. & Griffen, B.D. (2010). Early warning signals of extinction in
deteriorating environments. Nature, 467, 456–459.
FAO. (2020). FAO. 2020. Global Forest Resources Assessment 2020 – Key
findings. Rome. https://doi.org/10.4060/ca8753en.
Fattahi, B. & Ildoromi, A.R. (2011). Effect of Some Environmental Factors on
Plant Species Diversity in the Mountainous Grasslands (Case Study:
Hamedan - Iran). Int. J. Nat. Resour. Mar. Sci., 1, 45–52.
Filippa, G., Cremonese, E., Migliavacca, M., Galvagno, M., Forkel, M.,
Wingate, L., et al. (2016). Phenopix: AR package for image-based vegetation
phenology. Agric. For. Meteorol., 220, 141–150.
FIRMS. (2018). Fire Information for Resource Management System.
Available at: https://earthdata.nasa.gov/earth-observation-data/near-real-
time/firms/c6-mcd14dl. Last accessed 16 December 2018.

61
Friedl, M. & Sulla-Menashe, D. (2019). MCD12Q1 MODIS/Terra+Aqua Land
Cover Type Yearly L3 Global 500m SIN Grid V006 [Data set], NASA
EOSDIS L. Process. DAAC..
Gao, B.-C. (1996). NDWI—A normalized difference water index for remote
sensing of vegetation liquid water from space. Remote Sens. Environ., 58,
257–266.
Gatebe, C.K., Ichoku, C.M., Poudyal, R., Román, M.O. & Wilcox, E. (2014).
Surface albedo darkening from wildfires in northern sub-Saharan Africa.
Environ. Res. Lett., 9, 65003.
GCOS. (2003). Second report on the adequacy of the global observing
systems for climate, Technical Report GCOS-82 (WMO/TD No 1143).
GCOS. (2006). Systematic observation requirements for satellite-based
products for climate, Technical Report GCOS-107, WMO/TD. 1338/.
Ghulam, A., Li, Z., Qin, Q. & Tong, Q. (2007). Exploration of the spectral
space based on vegetation index and albedo for surface drought estimation.
J. Appl. Remote Sens., 1, 13529.
Giorgi, F. & Lionello, P. (2008). Climate change projections for the
Mediterranean region. Glob. Planet Chang., 63, 90–104.
Gitas, I., Mitri, G., Veraverbeke, S. & Polychronaki, A. (2012). Advances in
remote sensing of post-fire vegetation recovery monitoring-a review. In:
Remote Sensing of Biomass-Principles and Applications. 1, p.334.
Gorelick, N., Hancher, M., Dixon, M., Ilyushchenko, S., Thau, D. & Moore, R.
(2017). Google Earth Engine: Planetary-scale geospatial analysis for
everyone. Remote Sens. Environ., 202, 18–27.
Govaerts, Y. & Lattanzio, A. (2008). Estimation of surface albedo increase
during the eighties Sahel drought from Meteosat observations. Glob. Planet.
Change, 64, 139–145.
Govaerts, Y. & Lattanzlo, A. (2007). Surface albedo response to Sahel
precipitation changes. Eos (Washington. DC)., 88, 25–26.
Gu, L., Post, W.M., Baldocchi, D.D., Black, T.A., Suyker, A.E., Verma, S.B., et
al. (2009). Characterizing the seasonal dynamics of plant community
photosynthesis across a range of vegetation types. In: Phenology of
ecosystem processes. Springer, pp. 35–58.
Hadi, H. (2018). Satellite optical remote sensing of forest canopy cover in
boreal and tropical biomes. ISBN: 978-952-60-8225-7.
Hassanzadeh, E., Zarghami, M. & Hassanzadeh, Y. (2012). Determining the
main factors in declining the Urmia Lake level by using system dynamics
modeling. Water Resour. Manag., 26, 129–145.
Hastie, T.J. & Tibshirani, R.J. (1990). Generalized additive models, volume
43 of Monographs on Statistics and Applied Probability. ISBN 978-041-23-
4390-2.
Hayes, D.J. & Sader, S.A. (2001). Comparison of change-detection
techniques for monitoring tropical forest clearing and vegetation regrowth in
a time series. Photogramm. Eng. Remote Sensing, 67, 1067–1075.
Heydari, M., Rostamy, A., Najafi, F. & Dey, D.C. (2017). Effect of fire severity
on physical and biochemical soil properties in Zagros oak (Quercus brantii
62
References

Lindl.) forests in Iran. J. For. Res., 28, 95–104.


Hijmans, R.J. & van Etten, J. (2012). raster: Geographic analysis and
modeling with raster data. R package version 2.0–12.
Hirota, M., Holmgren, M., Van Nes, E.H. & Scheffer, M. (2011). Global
resilience of tropical forest and savanna to critical transitions. Science (80-.
)., 334, 232–235.
Hoerling, M., Eischeid, J., Perlwitz, J., Quan, X., Zhang, T. & Pegion, P.
(2012). On the increased frequency of mediterranean drought. J. Clim., 25,
2146–2161.
Holling, C.S. (1973). Resilience and stability of ecological systems. Annu.
Rev. Ecol. Syst., 1–23.
Hosseini, A., Hosseini, S.M., Rahmani, A. & Azadfar, D. (2012). Effect of tree
mortality on structure of Brant’s oak (Quercus brantii) forests of Ilam
province of Iran. Iran. J. For. Poplar Res., 20, 565–577.
Hosseinzadeh Talaee, P., Tabari, H. & Sobhan Ardakani, S. (2014).
Hydrological drought in the west of Iran and possible association with large-
scale atmospheric circulation patterns. Hydrol. Process., 28, 764–773.
Hovi, A., Liang, J., Korhonen, L., Kobayashi, H. & Rautiainen, M. (2016).
Quantifying the missing link between forest albedo and productivity in the
boreal zone. Biogeosciences, 13, 6015.
Hovi, A., Lindberg, E., Lang, M., Arumäe, T., Peuhkurinen, J., Sirparanta, S.,
et al. (2019). Seasonal dynamics of albedo across European boreal forests:
Analysis of MODIS albedo and structural metrics from airborne LiDAR.
Remote Sens. Environ., 224, 365–381.
Hovi, A., Raitio, P. & Rautiainen, M. (2017). A spectral analysis of 25 boreal
tree species. Silva Fenn, 51. [7753].
IPCC. (2018). An IPCC Special Report on the impacts of global warming of
1.5°C. Intergov. Panel Clim. Chang. Sustainable Development, and Efforts
to Eradicate Poverty. https://www.ipcc.ch/sr15/ (Accessed
date:05.06.2020).
IPCC, I. (2007). IPCC fourth assessment report: climate change 2007.
Intergov. panel Clim. Chang. Cambridge Univ. Press. Cambridge, 213–252.
Jarvis, A., Reuter, H.I., Nelson, A. & Guevara, E. (2008). Hole-filled SRTM
for the globe Version 4. available from CGIAR-CSI SRTM 90m Database
(http//srtm. csi. cgiar. org).
Jin, Y. & Roy, D.P. (2005). Fire-induced albedo change and its radiative
forcing at the surface in northern Australia. Geophys. Res. Lett., 32, 1–4.
Jurc, D. & Ogris, N. (2006). First reported outbreak of charcoal disease
caused by Biscogniauxia mediterranea on Turkey oak in Slovenia. Plant
Pathol., 55, 299.
Kaniewski, D., Van Campo, E. & Weiss, H. (2012). Drought is a recurring
challenge in the Middle East. Proc. Natl. Acad. Sci., 109, 3862–3867.
Karlsson, K.-G., Riihela, A., Müller, R., Meirink, J.F., Sedlar, J., Stengel, M.,
et al. (2013). CLARA-A1: a cloud, albedo, and radiation dataset from 28 yr of
global AVHRR data. Atmos. Chem. Phys., 13, 5351–5367.
Kéfi, S., Guttal, V., Brock, W.A., Carpenter, S.R., Ellison, A.M., Livina, V.N.,

63
et al. (2014). Early warning signals of ecological transitions: Methods for
spatial patterns. PLoS One, 9, e92097.
Kendall, M.G. (1948). Rank correlation methods. No. 98. Harvard Book List
#99. ISBN: 978-1-4684-6685-0.
Key, C.H. & Benson, N.C. (2006). Landscape Assessment (LA). FIREMON:
Fire effects monitoring and inventory system. Gen. Tech. Rep. RMRS-GTR-
164-CD.
Knyazikhin, Y., Martonchik, J. V, Myneni, R.B., Diner, D.J. & Running, S.W.
(1998). Synergistic algorithm for estimating vegetation canopy leaf area
index and fraction of absorbed photosynthetically active radiation from
MODIS and MISR data. J. Geophys. Res. Atmos., 103, 32257–32275.
Kubelka, P. & Munk, F. (1931). An article on optics of paint layers. Z. Tech.
Phys, 12.
Kuusinen, N. (2014). Boreal forest albedo and its spatial and temporal
variation. Diss. For. ISBN: 978-951-651-448-5.
Kuusinen, N., Kolari, P., Levula, J., Porcar-Castell, A., Stenberg, P. &
Berninger, F. (2012). Seasonal variation in boreal pine forest albedo and
effects of canopy snow on forest reflectance. Agric. For. Meteorol., 164, 53–
60.
Kuusinen, N., Lukeš, P., Stenberg, P., Levula, J., Nikinmaa, E. & Berninger,
F. (2014). Measured and modelled albedos in Finnish boreal forest stands of
different species, structure and understory. Ecol. Modell., 284, 10–18.
Kuusinen, N., Stenberg, P., Korhonen, L., Rautiainen, M. & Tomppo, E.
(2016). Structural factors driving boreal forest albedo in Finland. Remote
Sens. Environ., 175, 43–51.
Kuusk, A. & Nilson, T. (2000). A directional multispectral forest reflectance
model. Remote Sens. Environ., 72, 244–252.
Lamigueiro, O. & Hijmans, R. (2018). rasterVis. R package version 0.45.
(Accessed date: 05.06.2020).
Lee, X., Goulden, M.L., Hollinger, D.Y., Barr, A., Black, T.A., Bohrer, G., et al.
(2011). Observed increase in local cooling effect of deforestation at higher
latitudes. Nature, 479, 384–387.
Lelieveld, J., Hadjinicolaou, P., Kostopoulou, E., Chenoweth, J., El Maayar,
M., Giannakopoulos, C., et al. (2012). Climate change and impacts in the
Eastern Mediterranean and the Middle East. Clim. Change, 114, 667–687.
Lelieveld, J., Proestos, Y., Hadjinicolaou, P., Tanarhte, M., Tyrlis, E. & Zittis,
G. (2016). Strongly increasing heat extremes in the Middle East and North
Africa (MENA) in the 21st century. Clim. Change, 137, 245–260.
Lewis, P., Brockmann, C. & Muller, J.P. (2013). GlobAlbedo: Algorithm
theoretical basis document V4-12. In: Technical Report. 295pp.
Li, Y., Zhao, M., Motesharrei, S., Mu, Q., Kalnay, E. & Li, S. (2015). Local
cooling and warming effects of forests based on satellite observations. Nat.
Commun., 6, 6603.
Li, Z. & Garand, L. (1994). Estimation of surface albedo from space: A
parameterization for global application. J. Geophys. Res. Atmos., 99, 8335–
8350.
64
References

Liang, S. (2000). Numerical experiments on the spatial scaling of land


surface albedo and leaf area index. Remote Sens. Rev., 19, 225–242.
Liang, S. (2005). Quantitative remote sensing of land surfaces. John Wiley
& Sons. ISBN: 978-0-471-72371-4.
Liang, S., Zhao, X., Liu, S., Yuan, W., Cheng, X., Xiao, Z., et al. (2013). A
long-term Global LAnd Surface Satellite (GLASS) data-set for environmental
studies. Int. J. Digit. Earth, 6, 5–33.
Linaldeddu, B.T., Sirca, C., Spano, D. & Franceschini, A. (2011). Variation of
endophytic cork oak‐associated fungal communities in relation to plant
health and water stress. For. Pathol., 41, 193–201.
Liu, Y., Kumar, M., Katul, G.G. & Porporato, A. (2019). Reduced resilience as
an early warning signal of forest mortality. Nat. Clim. Chang., 9, 880–885.
Lu, Y., Nakicenovic, N., Visbeck, M. & Stevance, A.-S. (2015). Policy: five
priorities for the UN sustainable development goals. Nature, 520, 432–433.
Lukeš, P., Rautiainen, M., Manninen, T., Stenberg, P. & Mottus, M. (2014).
Geographical gradients in boreal forest albedo and structure in Finland.
Remote Sens. Environ., 152, 526–535.
Lukeš, P., Stenberg, P., Mõttus, M., Manninen, T. & Rautiainen, M. (2016).
Multidecadal analysis of forest growth and albedo in boreal Finland. Int. J.
Appl. earth Obs. Geoinf., 52, 296–305.
Lukeš, P., Stenberg, P. & Rautiainen, M. (2013a). Relationship between forest
density and albedo in the boreal zone. Ecol. Modell., 261–262, 74–79.
Lukeš, P., Stenberg, P., Rautiainen, M., Mottus, M. & Vanhatalo, K.M.
(2013b). Optical properties of leaves and needles for boreal tree species in
Europe. Remote Sens. Lett., 4, 667–676.
Lyons, E.A., Jin, Y.F. & Randerson, J.T. (2008). Changes in surface albedo
after fire in boreal forest ecosystems of interior Alaska assessed using MODIS
satellite observations. J. Geophys. Res., 113.
May, R.M. (1977). Thresholds and breakpoints in ecosystems with a
multiplicity of stable states. Nature, 269, 471–477.
May, R.M., Levin, S.A. & Sugihara, G. (2008). Ecology for bankers. Nature,
451, 893–894.
McDowell, N.G., Allen, C.D., Anderson-Teixeira, K., Aukema, B.H., Bond-
Lamberty, B., Chini, L., et al. (2020). Pervasive shifts in forest dynamics in a
changing world. Science (80-. )., 368.
Naimi, B. (2016). rts: Raster time series analysis. R Packag. version, 0–1.
(Accessed date: 05.06.2020).
Naimi, B., Hamm, N.A.S., Groen, T.A., Skidmore, A.K. & Toxopeus, A.G.
(2014). Where is positional uncertainty a problem for species distribution
modelling? Ecography (Cop.)., 37, 191–203.
Naimi, B., Hamm, N.A.S., Groen, T.A., Skidmore, A.K., Toxopeus, A.G. &
Alibakhshi, S. (2019). ELSA: Entropy-based local indicator of spatial
association. Spat. Stat., 29, 66–88.
Naudts, K., Chen, Y., McGrath, M.J., Ryder, J., Valade, A., Otto, J., et al.
(2016). Europe’s forest management did not mitigate climate warming.
Science (80-. )., 351, 597–600.

65
Ni, W. & Woodcock, C.E. (2000). Effect of canopy structure and the presence
of snow on the albedo of boreal conifer forests. J. Geophys. Res. Atmos., 105,
11879–11888.
Nicholson, E., Mace, G.M., Armsworth, P.R., Atkinson, G., Buckle, S.,
Clements, T., et al. (2009). Priority research areas for ecosystem services in a
changing world. J. Appl. Ecol., 46, 1139–1144.
Nijp, J.J., Temme, A.J.A.M., van Voorn, G.A.K., Kooistra, L., Hengeveld,
G.M., Soons, M.B., et al. (2019). Spatial early warning signals for impending
regime shifts: A practical framework for application in real‐world landscapes.
Glob. Chang. Biol., 25, 1905–1921.
O’Halloran, T.L., Law, B.E., Goulden, M.L., Wang, Z., Barr, J.G., Schaaf, C.,
et al. (2012). Radiative forcing of natural forest disturbances. Glob. Chang.
Biol., 18, 555–565.
Olson, D.M., Dinerstein, E., Wikramanayake, E.D., Burgess, N.D., Powell,
G.V.N., Underwood, E.C., et al. (2001). Terrestrial Ecoregions of the World:
A New Map of Life on EarthA new global map of terrestrial ecoregions
provides an innovative tool for conserving biodiversity. Bioscience, 51, 933–
938.
Pekel, J.-F., Cottam, A., Gorelick, N. & Belward, A.S. (2016). High-resolution
mapping of global surface water and its long-term changes. Nature, 540, 418.
Petus, C., Lewis, M. & White, D. (2013). Monitoring temporal dynamics of
Great Artesian Basin wetland vegetation, Australia, using MODIS NDVI.
Ecol. Indic., 34, 41–52.
Pitman, A.J., De Noblet-Ducoudré, N., Cruz, F.T., Davin, E.L., Bonan, G.B.,
Brovkin, V., et al. (2009). Uncertainties in climate responses to past land
cover change: First results from the LUCID intercomparison study. Geophys.
Res. Lett., 36.
Planque, C., Carrer, D. & Roujean, J.-L. (2017). Analysis of MODIS albedo
changes over steady woody covers in France during the period of 2001–2013.
Remote Sens. Environ., 191, 13–29.
Popp, C., Wang, P., Brunner, D., Stammes, P., Zhou, Y. & Grzegorski, M.
(2011). MERIS albedo climatology for FRESCO+ O2 A-band cloud retrieval.
Atmos. Meas. Tech., 4, 463–483.
Pourreza, M., Hosseini, S.M., Akbar, A., Sinegani, S., Matinizadeh, M., Dick,
W.A., et al. (2014). Soil microbial activity in response to fire severity in
Zagros oak ( Quercus ...
https://www.researchgate.net/publication/259089729_Soil_microbial_acti
... Article Soil microbial activity in response to fire severity in Zagros.
Geoderma, 213, 2014.
R Core Team. (2012). R: A Language and Environment for Statistical
Computing. R Foundation for Statistical Computing. R Foundation for
Statistical Computing.
http://finzi.psych.upenn.edu/R/library/dplR/doc/intro-dplR.pdf. (Accessed
date: 05.06.2020).
Ratajczak, Z., Carpenter, S.R., Ives, A.R., Kucharik, C.J., Ramiadantsoa, T.,
Stegner, M.A., et al. (2018). Abrupt change in ecological systems: inference
and diagnosis. Trends Ecol. Evol., 33, 513–526.
Rautiainen, M., Heiskanen, J. & Korhonen, L. (2012). Seasonal changes in
66
References

canopy leaf area index and moDis vegetation products for a boreal forest site
in central Finland. Boreal Environ. Res., 17.
Rautiainen, M. & Lukeš, P. (2015). Spectral contribution of understory to
forest reflectance in a boreal site: an analysis of EO-1 Hyperion data. Remote
Sens. Environ., 171, 98–104.
Rautiainen, M., Mõttus, M., Heiskanen, J., Akujärvi, A., Majasalmi, T. &
Stenberg, P. (2011). Seasonal reflectance dynamics of common understory
types in a northern European boreal forest. Remote Sens. Environ., 115,
3020–3028.
Reygadas, Y., Jensen, J.L.R. & Moisen, G.G. (2019). Forest Degradation
Assessment Based on Trend Analysis of MODIS-Leaf Area Index: A Case
Study in Mexico. Remote Sens., 11, 2503.
Riihelä, A. (2013). The Surface Albedo of the Arctic from Spaceborne Optical
Imagers: Retrieval and Validation. Finnish Meteorological Institute.
Roberts, D.A., Ustin, S.L., Ogunjemiyo, S., Greenberg, J., Dobrowski, S.Z.,
Chen, J., et al. (2004). Spectral and structural measures of northwest forest
vegetation at leaf to landscape scales. Ecosystems, 7, 545–562.
Rogers, B.M., Solvik, K., Hogg, E.H., Ju, J., Masek, J.G., Michaelian, M., et
al. (2018). Detecting early warning signals of tree mortality in boreal North
America using multiscale satellite data. Glob. Chang. Biol., 24, 2284–2304.
Rokni, K., Ahmad, A., Selamat, A. & Hazini, S. (2014). Water feature
extraction and change detection using multitemporal landsat imagery.
Remote Sens., 6, 4173–4189.
Sagan, C., Toon, O.B. & Pollack, J.B. (1979). Anthropogenic albedo changes
and the earth’s climate. Science (80-. )., 206, 1363–1368.
Schaaf, C. & Wang, Z. (2015). MCD43A3 MODIS/Terra+ Aqua BRDF/Albedo
Daily L3 Global—500 m V006. LP DAAC. [Data set]. (Accessed date:
05.06.2020).
Schaaf, C.B., Gao, F., Strahler, A.H., Lucht, W., Li, X., Tsang, T., et al.
(2002). First operational BRDF, albedo nadir reflectance products from
MODIS. Remote Sens. Environ., 83, 135–148.
Schaepman-Strub, G., Schaepman, M.E., Painter, T.H., Dangel, S. &
Martonchik, J. V. (2006). Reflectance quantities in optical remote sensing—
Definitions and case studies. Remote Sens. Environ., 103, 27–42.
Scheffer, M. (2009). Critical transitions in nature and society. Choice Rev.
Online. Princeton University Press. ISBN: 978-069-112204-5
Scheffer, M. (2010). Complex systems: Foreseeing tipping points. Nature.
Scheffer, M., Carpenter, S., Foley, J.A., Folke, C. & Walker, B. (2001).
Catastrophic shifts in ecosystems. Nature, 413, 591–596.
Scheffer, M., Carpenter, S.R., Lenton, T.M., Bascompte, J., Brock, W., Dakos,
V., et al. (2012). Anticipating critical transitions. Science (80-. ). American
Association for the Advancement of Science.
Schwaiger, H.P. & Bird, D.N. (2010). Integration of albedo effects caused by
land use change into the climate balance: Should we still account in
greenhouse gas units? For. Ecol. Manage., 260, 278–286.
Sellers, P.J. (1987). Canopy reflectance, photosynthesis, and transpiration, II.

67
The role of biophysics in the linearity of their interdependence. Remote Sens.
Environ., 21, 143–183.
Sen, P.K. (1968). Estimates of the regression coefficient based on Kendall’s
tau. J. Am. Stat. Assoc., 63, 1379–1389.
Sentinel, E.S.A. (2013). User Handbook. ESA Paris, Fr.
Sexton, J.O., Song, X.-P., Feng, M., Noojipady, P., Anand, A., Huang, C., et
al. (2013). Global, 30-m resolution continuous fields of tree cover: Landsat-
based rescaling of MODIS vegetation continuous fields with lidar-based
estimates of error. Int. J. Digit. Earth, 6, 427–448.
Shadkam, S., Ludwig, F., van Oel, P., Kirmit, Ç. & Kabat, P. (2016). Impacts
of climate change and water resources development on the declining inflow
into Iran’s Urmia Lake. J. Great Lakes Res., 42, 942–952.
Sims, D.A. & Gamon, J.A. (2002). Relationships between leaf pigment
content and spectral reflectance across a wide range of species, leaf
structures and developmental stages. Remote Sens. Environ., 81, 337–354.
Solano, R., Didan, K., Jacobson, A. & Huete, A. (2010). MODIS vegetation
index user’s guide (MOD13 series). Veg. Index Phenol. Lab, Univ. Arizona,
1–38.
Sonnenschein, R., Kuemmerle, T., Udelhoven, T., Stellmes, M. & Hostert, P.
(2011). Differences in Landsat-based trend analyses in drylands due to the
choice of vegetation estimate. Remote Sens. Environ., 115, 1408–1420.
Stenberg, P., Lukeš, P., Rautiainen, M. & Manninen, T. (2013). A new
approach for simulating forest albedo based on spectral invariants. Remote
Sens. Environ., 137, 12–16.
Stocker, T.F., Qin, D., Plattner, G.-K., Tignor, M., Allen, S.K., Boschung, J., et
al. (2013). Climate change 2013: The physical science basis. Contrib. Work.
Gr. I to fifth Assess. Rep. Intergov. panel Clim. Chang., 1535.
Strahler, A.H., Muller, J.P., Lucht, W., Schaaf, C., Tsang, T., Gao, F., et al.
(1999). MODIS BRDF/albedo product: algorithm theoretical basis document
version 5.0. MODIS Doc., 23, 42–47.
Su, Z. (2002). The Surface Energy Balance System (SEBS) for estimation of
turbulent heat fluxes. Hydrol. earth Syst. Sci., 6, 85–99.
Sun, G., Noormets, A., Gavazzi, M.J., McNulty, S.G., Chen, J., Domec, J.-C.,
et al. (2010). Energy and water balance of two contrasting loblolly pine
plantations on the lower coastal plain of North Carolina, USA. For. Ecol.
Manage., 259, 1299–1310.
TAMGA, D.E.K. (2019). Critical transition in ecosystems: comparing the
effect of detrending techniques on early warning signals.
https://www.researchgate.net/profile/Dan_Kanmegne/publication/3377171
67_CRITICAL_TRANSITION_IN_ECOSYSTEMS_COMPARING_THE_EF
FECT_OF_DETRENDING_TECHNIQUES_ON_EARLY_WARNING_SIGN
ALS/links/5de67f764585159aa45f53ac/CRITICAL-TRANSITION-IN-
ECOSYSTEMS-COMPARING-THE-EFFECT-OF-DETRENDING-
TECHNIQUES-ON-EARLY-WARNING-SIGNALS.pdf. (accessed date:
05.06.2020).
Team, Q.D. (1991). QGIS Geographic Information System. Open Source
Geospatial Foundation Project. (Accessed date: 05.06.2020).
68
References

Tian, L., Chen, J. & Shao, C. (2018). Interdependent Dynamics of LAI-Albedo


across the Roofing Landscapes: Mongolian and Tibetan Plateaus. Remote
Sens., 10, 1159.
Trigo, R.M., Gouveia, C.M. & Barriopedro, D. (2010). The intense 2007-2009
drought in the Fertile Crescent: Impacts and associated atmospheric
circulation. Agric. For. Meteorol., 150, 1245–1257.
Tucker, C.J. (1978). Red and photographic infrared linear combinations for
monitoring vegetation. NASA-TM-79620. Accession number: 78N32525
Turner, M.G., Calder, W.J., Cumming, G.S., Hughes, T.P., Jentsch, A.,
LaDeau, S.L., et al. (2020). Climate change, ecosystems and abrupt change:
science priorities. Philos. Trans. R. Soc. B, 375, 20190105.
Vanderhoof, M., Williams, C.A., Shuai, Y., Jarvis, D., Kulakowski, D. &
Masek, J. (2014). Albedo-induced radiative forcing from mountain pine
beetle outbreaks in forests, south-central Rocky Mountains: magnitude,
persistence, and relation to outbreak severity. Biogeosciences, 11, 563–575.
Vannini, A., Lucero, G., Anselmi, N. & Vettraino, A.M. (2009). Response of
endophytic Biscogniauxia mediterranea to variation in leaf water potential of
Quercus cerris. For. Pathol., 39, 8–14.
Veraverbeke, S. (2010). Fire-induced changes in vegetation, albedo and land
surface temperature assessed with MODIS. In: … : Remote sensing for ….
European Association of Remote Sensing Laboratories (EARSeL), pp. 431–
438.
Veraverbeke, S., Lhermitte, S., Verstraeten, W.W. & Goossens, R. (2010). The
temporal dimension of differenced Normalized Burn Ratio (dNBR) fire/burn
severity studies: The case of the large 2007 Peloponnese wildfires in Greece.
Remote Sens. Environ., 114, 2548–2563.
Verbesselt, J. & Hyndman, R. (2010). Detecting trend and seasonal changes
in satellite image time series. Remote Sens. …, 114, 106–115.
Verbesselt, J., Hyndman, R., Zeileis, A. & Culvenor, D. (2010). Phenological
change detection while accounting for abrupt and gradual trends in satellite
image time series. Remote Sens. Environ., 114, 2970–2980.
Verbesselt, J., Zeileis, A. & Herold, M. (2012a). Near real-time disturbance
detection using satellite image time series. Remote Sens. Environ., 123, 98–
108.
Verbesselt, J., Zeileis, A. & Herold, M. (2012b). Near real-time disturbance
detection using satellite image time series. Remote Sens. Environ. 123,
pp.98-108.
Vermote, E. (2015). MOD09A1 MODIS/terra surface reflectance 8-day L3
global 500m SIN grid V006. NASA EOSDIS L. Process. DAAC, 10.
Vescovi, F.D. (2017). Index based mapping.
https://www.academia.edu/download/54358212/DroughtWaterIndices_vIP
P.pdf. (Accessed date: 05.06.2020).
Vicente-Serrano, S.M., Quiring, S.M., Peña-Gallardo, M., Yuan, S. &
Domínguez-Castro, F. (2020). A review of environmental droughts:
Increased risk under global warming? Earth-Science Rev., 201, 102953.
Voss, K.A., Famiglietti, J.S., Lo, M., De Linage, C., Rodell, M. & Swenson,
S.C. (2013). Groundwater depletion in the Middle East from GRACE with

69
implications for transboundary water management in the Tigris-Euphrates-
Western Iran region. Water Resour. Res., 49, 904–914.
Walters, C. & Kitchell, J.F. (2001). Cultivation/depensation effects on
juvenile survival and recruitment: implications for the theory of fishing. Can.
J. Fish. Aquat. Sci., 58, 39–50.
Wang, Q., Adiku, S., Tenhunen, J. & Granier, A. (2004). On the relationship
of NDVI with leaf area index in a deciduous forest site. Remote Sens.
Environ., 94, 244–255.
Wang, S. (2005). Dynamics of surface albedo of a boreal forest and its
simulation. Ecol. Modell., 183, 477–494.
Watson, D.J. (1947). Comparative physiological studies on the growth of field
crops: II. The effect of varying nutrient supply on net assimilation rate and
leaf area. Ann. Bot., 11, 375–407.
van der Werf, G.R., Randerson, J.T., Giglio, L., van Leeuwen, T.T., Chen, Y.,
Rogers, B.M., et al. (2017). Global fire emissions estimates during
1997&amp;amp;ndash;2015. Earth Syst. Sci. Data Discuss., 9, 1–43.
Wickham, H. (2016). ggplot2: elegant graphics for data analysis. Springer.
(Accessed date: 05.06.2020).
Wissel, C. (1984). A universal law of the characteristic return time near
thresholds. Oecologia, 65, 101–107.
Wolfe, R.E., Nishihama, M., Fleig, A.J., Kuyper, J.A., Roy, D.P., Storey, J.C.,
et al. (2002). Achieving sub-pixel geolocation accuracy in support of MODIS
land science. Remote Sens. Environ., 83, 31–49.
Wood, S.N. (2000). Modelling and smoothing parameter estimation with
multiple quadratic penalties. J. R. Stat. Soc. Ser. B (Statistical Methodol., 62,
413–428.
Wood, S.N. (2011). Fast stable restricted maximum likelihood and marginal
likelihood estimation of semiparametric generalized linear models. J. R. Stat.
Soc. Ser. B (Statistical Methodol., 73, 3–36.
Wood, S.N. (2017). Generalized additive models: an introduction with R.
Chapman and Hall/CRC.
Woolley, J.T. (1971). Reflectance and transmittance of light by leaves. Plant
Physiol., 47, 656–662.
Wu, C., Gonsamo, A., Gough, C.M., Chen, J.M. & Xu, S. (2014). Modeling
growing season phenology in North American forests using seasonal mean
vegetation indices from MODIS. Remote Sens. Environ., 147, 79–88.
Xu, H. (2006). Modification of normalised difference water index (NDWI) to
enhance open water features in remotely sensed imagery. Int. J. Remote
Sens., 27, 3025–3033.
Xue, X., Liu, H., Mu, X. & Liu, J. (2014). Trajectory-based detection of urban
expansion using Landsat time series. Int. J. Remote Sens., 35, 1450–1465.
Yue, S. & Hashino, M. (2003). Long term trends of annual and monthly
precipitation in Japan 1. JAWRA J. Am. Water Resour. Assoc., 39, 587–596.
Zhang, Y., Tian, Y., Knyazikhin, Y., Martonchik, J. V, Diner, D.J., Leroy, M.,
et al. (2000). Prototyping of MISR LAI and FPAR algorithm with POLDER
70
References

data over Africa. IEEE Trans. Geosci. Remote Sens., 38, 2402–2418.
Zhao, B., Yan, Y., Guo, H., He, M., Gu, Y. & Li, B. (2009). Monitoring rapid
vegetation succession in estuarine wetland using time series MODIS-based
indicators: An application in the Yangtze River Delta area. Ecol. Indic., 9,
346–356.
Zhao, K. & Jackson, R.B. (2014). Biophysical forcings of land‐use changes
from potential forestry activities in North America. Ecol. Monogr., 84, 329–
353.
Zheng, X. & Basher, R.E. (1999). Structural time series models and trend
detection in global and regional temperature series. J. Clim., 12, 2347–2358.
Zhu, Z. (2017). Change detection using landsat time series: A review of
frequencies, preprocessing, algorithms, and applications. ISPRS J.
Photogramm. Remote Sens., 130, 370–384.

71
The earth is under unprecedented pressure, which is refl ected in rapid

Aalto-DD 183/2020
ecosystem changes around the globe. Over just the past three decades,
the earth has lost over 178 million hectares of its forests. The rapidly
growing evidence of the loss of resilience in ecosystems (i.e., recovery
from disturbances slows down) due to climate change has become a
global concern. Our limited knowledge of ecosystem dynamics and their
key parameters, such as albedo, has also hindered our ability to manage
ecosystems appropriately.
The main aim of this dissertation is to contribute to elucidating
ecosystem dynamics by exploiting remotely sensed satellite data.
Additionally, this dissertation aims to explore the dynamics of albedo
(refl ectivity) in response to forest structure (forest density, tree cover,
and leaf area index) variations and forest disturbances (fi re and
drought).

BUSINESS +
ECONOMY
ISBN 978-952-64-0129-4 (pr int ed)
9HSTFMG*eabcje+

ART +
ISBN 978-952-64-0130-0 (pdf)
DESIGN +
ISSN 1799-4934 (pr int ed) ARCHITECTURE
ISSN 1799-4942 (pdf)
SCIENCE +
T E C H N O L O GY
Aalto University
C R O S S OV E R
School of Engineering
Department of Built Environment DOCTORAL
www.aalto.fi D I S S E R TA T I O N S

You might also like