ArticlePDF AvailableLiterature Review

Strategic roadmap to assess forest vulnerability under air pollution and climate change

Authors:

Abstract

Although it is an integral part of global change, most of the research addressing the effects of climate change on forests have overlooked the role of environmental pollution. Similarly , most studies investigating effects of air pollutants on forests have generally neglected impacts of climate change. We review the current knowledge on combined air pollution and climate change effects on global forest ecosystems and identify several key research priorities as a roadmap for the future. Specifically, we recommend 1) establishment of much denser array of monitoring sites, particularly in the South Hemisphere; 2) further integration of ground and satellite monitoring; 3) generation of flux-based standards and critical levels taking into account the sensitivity of dominant forest tree species; 4) long-term monitoring of N, S, P cycles and base cations deposition together at global scale; 5) intensification of experimental studies, addressing combined effects of different abiotic factors on forests by assuring a better representation of taxonomic and functional diversity across the ~ 73,000 tree species on Earth; 6) more experimental focus on phenomics and genomics; 7) improved knowledge on key processes regulating the dynamics of radionuclides in forest systems; and 8) development of models integrating air pollution and climate change data from long-term monitoring programs.
Glob Change Biol. 2022;00:1–24.
|
1wileyonlinelibrary.com/journal/gcb
Received: 6 October 2021 
|
Revised: 2 March 2022 
|
Accepted: 18 May 2022
DOI: 10.1111/gcb.16278
REVIEW
Strategic roadmap to assess forest vulnerability under air
pollution and climate change
Alessandra De Marco1| Pierre Sicard2| Zhaozhong Feng3| Evgenios Agathokleous3|
Rocio Alonso4| Valda Araminiene5| Algirdas Augustatis6| Ovidiu Badea7,8|
James C. Beasley9| Cristina Branquinho10| Viktor J. Bruckman11| Alessio Collalti12 |
Rakefet David- Schwartz13| Marisa Domingos14| Enzai Du15 | Hector Garcia Gomez4|
Shoji Hashimoto16| Yasutomo Hoshika17| Tamara Jakovljevic18| Steven McNulty19|
Elina Oksanen20| Yusef Omidi Khaniabadi21| Anne- Katrin Prescher22|
Costas J. Saitanis23| Hiroyuki Sase24| Andreas Schmitz25| Gabriele Voigt26|
Makoto Watanabe27| Michael D. Wood28| Mikhail V. Kozlov29 | Elena Paoletti16
1ENEA, CR Casaccia, SSPT- PVS, Rome, Italy
2ARGANS, Biot, France
3Key Laboratory of Agro- Meteorology of Jiangsu Province, School of Applied Meteorology, Nanjing University of Information Science & Technology, Nanjing, China
4Ecotoxicology of Air Pollution, CIEMAT, Madrid, Spain
5Lithuanian Research Centre for A griculture and Forestry, Kaunas, Lithuania
6Faculty of Forest Sciences and Ecology, Vytautas Magnus University, Kaunas, Lithuania
7“Marin Drăcea” National Institute for Research and Development in Forestry, Voluntari, Romania
8Faculty of Silviculture and Forest Engineering, “Transilvania” Universit y, Braşov, Romania
9Savannah River Ecology Laboratory and Warnell School of Forestry and Natural Resources, University of Georgia, Aiken, South Carolina, USA
10Centre for Ecology, Evolution and Environmental Changes, Faculdade de Ciências, Universidade de Lisboa, Lisbon, Portugal
11Commission for Interdisciplinary Ecological Studies, Austrian Academy of Sciences, Vienna, Austria
12Forest Modeling Lab., ISAFOM- CNR, Perugia, Italy
13Institute of Plant Sciences, AROVolcani Center, Rishon LeTsiyon, Israel
14Instituto de Botanica, Nucleo de Pesquisa em Ecologia, Sao Paulo, Brazil
15Faculty of Geographical Science, Beijing Normal University, Beijing, China
16Depar tment of Forest Soils, Forestry and Forest Products Research Institute, Tsukuba, Japan
17IRET- CNR, Sesto Fiorentino, Italy
18Croatian Forest Research Institute, Jastrebarsko, Croatia
19USDA Forest Service, Research Triangle Park, USA
20Depar tment of Environmental and Biological Sciences, University of Eastern Finland, Joensuu, Finland
21Department of Environmental Health Engineering, Industrial Medial and Health, Petroleum Industry Health Organization (PIHO), Ahvaz, Iran
22Thuenen Institute of Forest Ecosystems, Eberswalde, Germany
23Lab of Ecology and Environmental Science, Agricultural University of Athens, Athens, Greece
24Ecological Impact Research Department, Asia Center for Air Pollution Research (ACAP), Niigata, Japan
25State Agency for Nature, Environment and Consumer Protection of North Rhine- Westphalia, Recklinghausen, Germany
26r.e.m. Consulting, Perchtoldsdorf, Austria
27Institute of Agriculture, Tokyo University of Agriculture and Technology (TUAT), Fuchu, Japan
28School of Science, Engineering and Environment, University of Salford, Salford, UK
29Department of Biology, University of Turku, Turku, Finland
This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium,
provided the original work is properly cited.
© 2022 The Authors. Global Change Biology published by John Wiley & Sons Ltd.
2 
|
    DE MARCO et al.
1 | INTRODUC TION
Forests cover ~30% of the world's land surface, store 45% of terres-
trial carbon (Bonan, 2008), and are home to 80% of global terrestrial
biodiversity (IUCN, 2021). Sustainable socioeconomic development
depends on the proper management of natural resources, includ-
ing forest ecosystems (Badea et al., 2013). Air pollution and climate
change have major impacts on and complex interactions with forest
health and productivity (Augustaitis & Bytnerowicz, 2008; Kozlov
et al., 2009). For example, tropospheric ozone (O3), which is both
a phytotoxic gas and a radiative forcer (Myhre et al., 2013), and ni-
trogen deposition (Du & de Vries, 2018), which causes forest de-
cline due to acidification (Augustaitis et al., 2010) and changes in
the frequency and intensity of climatic extremes (e.g., heat waves,
rainfall, wind storms), may impact the structure, composition, and
functioning of terrestrial ecosystems. These impacts can directly in-
fluence carbon cycling and its feedback to the climate system (Frank
et al., 2015; Matyssek et al., 2012; Paoletti et al., 2007; Serengil
et al., 2011; Sicard et al., 2020).
The future of global forests is a subject of public and politi-
cal concern due to extensive forest degradation worldwide (Hao
et al., 2018; Liu et al., 2018). Recently, environmental pollution
was identified as one of the five main drivers of biodiversity loss
(European Commission, 2020). Although environmental pollution
is an integral part of global change (Dale et al., 2000), most of the
research addressing the biotic effects of climate change do not con-
sider this issue. Furthermore, most studies on both the distribution
of pollutants and the biotic effects of pollution have neglected the
issue of climate change (Sicard, Augustaitis, et al., 2016). As a result,
studies exploring the combined effects of air pollution and climate
change remain uncommon.
A Web of Science search conducted in June 2021 identified only
74 peer- reviewed articles containing the keywords “climat* and pol-
lut*” and “tree* or forest*” in the title, 59 of which were relevant re-
search papers (Figure S1): In all, 11 studies used modeling to explore
the combined effects of air pollution and climate, 27 studies were
based on observations of forest health in either spatial or tempo-
ral gradients of air pollution and climate, and only one reported the
outcomes of a field experiment. The low number of experimental
studies with factorial design involving both airborne pollutants and
climate is alarming because it hampers our ability to identify cause-
and- effect relationships as well as to decipher the mechanisms un-
derlying the combined or interactive effects of pollution and climate
on the health of individual trees and forest ecosystems. As a result,
the quality of our predictions of the combined effects of climate
change and air pollution on future forest health is uncertain. To re-
spond to this global challenge, here we critically review the current
knowledge (and gaps) on air pollution and climate interactions in for-
ests, identify key research priorities, and suggest a strategic road-
map for future studies.
Correspondence
Alessandra De Marco, ENEA Casaccia, Via
Anguillarese 301, 00123 Rome, Italy.
Email: alessandra.demarco@enea.it
Funding information
Academy of Finland, Grant/Award
Number: 276671, 311929 and 316182;
European Commission, Grant/Award
Number: LIFE15 ENV/IT/000183, LIFE19
ENV/FR/000086 and LIFE20 GIE/
IT/000091; JST SICORP, Grant/Award
Number: JPMJSC16HB; US Department of
Energy, Grant/Award Number:
DE- EM0005228
Abstract
Although it is an integral part of global change, most of the research addressing the
effects of climate change on forests have overlooked the role of environmental pollu-
tion. Similarly, most studies investigating the effects of air pollutants on forests have
generally neglected the impacts of climate change. We review the current knowledge
on combined air pollution and climate change effects on global forest ecosystems
and identify several key research priorities as a roadmap for the future. Specifically,
we recommend (1) the establishment of much denser array of monitoring sites, par-
ticularly in the South Hemisphere; (2) further integration of ground and satellite
monitoring; (3) generation of flux- based standards and critical levels taking into ac-
count the sensitivity of dominant forest tree species; (4) long- term monitoring of
N, S, P cycles and base cations deposition together at global scale; (5) intensification
of experimental studies, addressing the combined effects of different abiotic factors
on forests by assuring a better representation of taxonomic and functional diversity
across the ~73,000 tree species on Earth; (6) more experimental focus on phenomics
and genomics; (7) improved knowledge on key processes regulating the dynamics of
radionuclides in forest systems; and (8) development of models integrating air pollu-
tion and climate change data from long- term monitoring programs.
KEYWORDS
air pollution, climate change, forest ecosystem, forest nutrients, forest research roadmap,
forest vulnerability, radioactivity
   
|
 3
DE MARCO et al.
2 | ASSESSING AIR POLLUTION:
RESEARCH INFRASTRUCTURES AND
METHODOLOGIES FOR FOREST
MONITORING
Re giona l and nati ona l air qua lit y dir ec tive s and em issi ons contr ol po l-
icies (e.g., Japanese Air Pollution Control Act 1968/1970; European
Council Directive 2008/50/EC; United States Federal Register,
2015) led to the development of air quality monitoring stations.
Monitoring data are collated within national or regional databases,
such as the Acid Deposition Monitoring Network in East Asia, the
European Environment Agency Airbase system, and the Australia Air
Quality Network (AUSAQN; Schultz et al., 2017 ). Despite efforts to
moni tor air quality in Sout h America , the spat ial dist rib uti on of moni-
toring stations is still heterogeneous and insufficient to represent
the pollutant levels (Peláez et al., 2020).
Coordinated research networks of long- term experimental for-
est sites integrating monitoring and state- of- the- art methodological
and conceptual research to assess air pollution and global change
effects are not distributed in a way that represents all forest eco-
system types over the globe. Long- term forest monitoring and in-
frastructure networks are running regionally and worldwide, even
overlapping each other in their geographic expansions, and are likely
to further expand in the future. Here, we introduce some of the larg-
est networks of experimental forest sites, their research aims and
methodologies, and explore their capacities in view of the Supersite
definition (Mikkelsen et al., 2013).
International Long- Term Ecological Research (ILTER) is a “net-
work of networks” with research sites located in a wide array of
ecosystems aimed at developing a global understanding of envi-
ronmental change while also covering socioeconomical aspects
(known as LTSER). Expertise warrants the collection, management,
and analysis of spatiotemporally diverse datasets, such as DEIMS
(Drupal Ecological Information Management System), a central plat-
form providing information on sites and networks with ecological
long- term monitoring and experimentation at European and global
scales. Currently, ILTER encompasses 39 countries which together
operate more than 600 sites (Maass et al., 2016). Some sites main-
tain advanced continuous measurements, such as tower- based eddy
covariance assessments of CO2 and H2O fluxes. The ILTER network
includes the Ter res trial Eco sys tem Re search Netwo rk (TE RN), est ab-
lished in Australia, which provides a comprehensive metadata portal
containing information on continental scale datasets of measure-
ments describing fauna, flora, terrestrial ecosystems, ecological dy-
namics, land surfaces, soils, agricultural ecosystems, coasts, climate
observations and fluxes (Karan et al., 2016). Similarly, the Chinese
National Ecosystem Research Network (CNERN) is an integrated
platform of field stations supervised by various Chinese ministries.
CNERN represents a science and technology system that conducts
network observation and experimentation across China's ecosys-
tems, cutting across governmental departments, industrial sectors,
regions, and jurisdictions, and seeks to integrate observation equip-
ment and data resources and standardize research methods, tools,
and protocols. As a result, CNERN serves as a nexus for national
ecological research, promotes data sharing, and creates an educa-
tional center and collaborative base for ecological research. ILTER
networks are also present in Korea and Taiwan.
Another “network of regional networks” is represented by
FLUXNET, which is coordinating regional and global analyses con-
ducted at micrometeorological tower sites (eddy covariance) to
investigate the exchanges of carbon dioxide (CO2), water vapor,
and energy between terrestrial ecosystems and the atmosphere
(Pastorello et al., 2020). FLUXNET is divided into regional networks,
for example, the European Integrated Carbon Observation System
Research Infrastructure (ICOS RI) with more than 100 measuring
stations including 32 forest stations. In 2021, more than 800 sites
worldwide were operated on a long- term and continuous basis
within this network. Habitats included in this monitoring framework
include temperate conifer and broadleaf (deciduous and evergreen)
forests, tropical and boreal forests, crops, grasslands, chaparral,
wetlands, and tundra.
In Europe, the International Co- operative Programme on
Assessment and Monitoring of Air Pollution Effects was launched
in 1985 under the United Nations Convention on Long- Range
Transboundary Air Pollution (CLRTAP), with several units including
ICP Forests (Michel et al., 2018), ICP Vegetation, ICP Modelling and
Mapping, and ICP Integrated Monitoring (Lundin & Forsius, 2004).
Networks of monitoring stations are established within this frame-
work that continuously assess ecosystem responses to air pollution
and develop the associated modeling and assessment methods
(Forsius et al., 2021). ICP Forests currently monitor forest conditions
in Europe at two intensities: Level I is based on around 6000 ob-
servation plots within a systematic transnational grid of 16 × 16 km2.
Level II comprises around 800 plots in selected forest ecosystems
for clarifying cause– effect relationships, and also assesses foliar and
soil chemistry, tree growth, and conditions of ground vegetation.
Approximately 41 sites, depending on the parameters, also monitor
ambient air quality and meteorology.
ForestGEO is a global network of scientists and forest research
sites dedicated to advancing long- term study of the world's forests,
dedicated to the study of tropical and temperate forest function and
diversity. The multi- institutional network comprises 73 forest re-
search sites across the Americas, Africa, Asia, Europe, and Oceania.
ForestGEO monitors the growth and survival of approximately 6 mil-
lion trees and nearly 13,000 species that occur in the forest research
sites. This network also supports initiatives to monitor attributes
such as climate, carbon flux, vertebrates, insects, and soil microor-
ganisms. ForestGEO increases scientific understanding about the
potential effects of climate change on ecosystems, which is a priority
of the US Climate Change Science Program and highlighted by the
Intergovernmental Panel on Climate Change (IPCC) Working Group
II. Because of ForestGEO's extensive biological monitoring, unique
databases, and the partners' expertise, it promises to enhance soci-
ety's ability to evaluate and respond to the impacts of global climate
change. To date, unfortunately, the distribution of forest monitor-
ing sites within ForestGEO appears non- homogeneous (Figure 1).
4 
|
    DE MARCO et al.
Indeed, boreal and tropical forest s are less represented among mon-
itoring sites and there is a disproportionate number of monitoring
sites in the Northern Hemisphere (NH), particularly in Europe, and
fewer sites in the Southern Hemisphere (SH).
The aforementioned monitoring networks may benefit from
data derived through remote sensing measurements (Lechner
et al., 2020). Remotely sensed imagery provides a synoptic view, and
is potentially available everywhere at a large range of spatial and
temporal scales with a high degree of homogeneity. Furthermore,
remote imagery provides digital images that can easily be integrated
with other spatial datasets in a geographic information system, and
per unit area remote sensing is an inexpensive way to acquire data.
The most used remote sensing sensors for assessing and monitoring
forest conditions are those on- board satellites, followed by airborne
(including Unmanned Aerial Vehicles) and terrestrial systems, or a
combination of these platforms (Torres et al., 2021). Previous studies
have demonstrated the utility of optical remote sensing for assessing
a variety of forest health indices, and are commonly used in forest
monitoring activities (Curran et al., 1992; Huang et al., 2019 ; Parent
& Verbyla, 2010 ). Landsat satellite images are still the most widely
used Earth Observation (EO) data in forest health studies (Torres
et al., 2021), which provide continuous time series data from the
1970s (i.e., Landsat 1 mission) until today (i.e., Landsat 8). Access to
Landsat images has been free since 2008, and the recently launched
Landsat 9 (September 2021) will be publicly available in early 2022.
In addition to Landsat imagery, imagery from sentinel missions
from the European Space Agency is also particularly important for
forest monitoring because of their high spatial and temporal reso-
lution. Furthermore, the availability of both active (Sentinel- 1) and
passive (Sentinel- 2) sensors might increase the precision of previ-
ous analytical methods that rely primarily on optical reflectance
indices. Similarly, forest health monitoring studies are increasingly
using Synthetic Aperture Radar (SAR) sensors. For example, C- band
data are sensitive to variations of Leaf Area Index, which are con-
nected to defoliation and hence forest status (Manninen et al., 2003;
Stankevich et al., 2017). SAR sensors are advantageous not just
because of their sensitivity to forest structural changes (Dobson
et al., 1992; Harrell et al., 1995; Le Toan et al., 1992), but also be-
cause of their ability to monitor the water content of the tree canopy
(Dobson et al., 199 2; Harrell et al., 1995; Le Toan et al., 199 2).
Specific remote sensing techniques that merge different spatial,
spectral, radiometric, and temporal resolutions are being increas-
ingly used to reduce data gaps and to characterize forest ecosystems
(Lausch et al., 2018). For example, Rogers et al. (2018) demonstrated
the potential of derived products based on Landsat, Advanced
Very High- Resolution Radiometer (AVHRR), and MODIS (Moderate
Resolution Imaging Spectroradiometer) data to detect early signals
of tree mortality. Modeling various biophysical indicators based on
aerial or ground- based LiDAR data can further expand the portfo-
lio of remote sensing- derived data, or at the very least allow their
validation in a more efficient manner than by means of traditional
monitoring and inventory. In this regard, a fusion of satellite spectral
data (e.g., Sentinel- 2) and LiDAR data (Global Ecosystem Dynamics
Investigations) could be the next step for global drought- induced
tree mortality assessment (Huang et al., 2019). More recently, the
Copernicus air- pollution monitoring satellite dedicated to trace gas-
ses assessment, such as O3, NO2, SO2, formaldehyde (HCHO), CO,
and CH4 (Sentinel- 5— Precursor/TROPOMI; Inness et al., 2019),
has been used for tracking pollution events and pollution sources
(Mesas- Carrascosa et al., 2020). By merging classical monitoring
techniques and state- of- the- art remote sensing, long- term studies
are facilitated (Tănase et al., 2019). Remote sensing use should be
FIGURE 1 Distribution of the most relevant monitoring network over the forested areas of the globe.
   
|
 5
DE MARCO et al.
expand ed to vulnerable regio ns or ecosystem ty pes whi ch need spe-
cial protection from climate change and air pollution.
Highly instrumented forest research infrastructures (supersites)
provide long- term data series and promote integration of research
communities in a transcontinental collaboration network (Fischer
et al., 2011). For these supersites, the use of forest inventory data
together with remote sensing and EO data can provide valuable in-
formation on forest condition (Hartmann et al., 2018). As new forest
change detection algorithms based on EO sensors are developed,
they can be validated using data from long- term monitoring net-
works (Rodman et al., 2021).
To understand climate change and weather extremes, it is im-
portant to have observations of the Earth system going back as far
as possible in time. Reanalysis combines past short- range weather
forecasts with observations through data assimilation (Uppala
et al., 2005). The process mimics the production of day- to- day
weather forecasts. Reanalyses are usually produced at lower resolu-
tion than current weather forecasts, and they use the same modern-
data assimilation system and forecasting model throughout the
reanalysis period. The latest European Centre for Medium- Range
Weather Forecasts (ECMWF) reanalyses are produced through the
EU- funded Copernicus Climate Change Service (C3S). Forecasts are
freely available through the C3S Climate Data Store. The most recent
ECMWF reanalysis dataset is the ER A5 Back Extension, providing
data from 1950 to 1978. The Copernicus Atmosphere Monitoring
Service (CAMS) provides continuous data and information on atmo-
spheric composition. The service describes the current situation,
forecasts the situation a few days ahead, and analyses consistently
retrospective data records for recent years. CAMS supports many
applications in a variety of domains including health, environmen-
tal monitoring, renewable energies, meteorology, and climatology.
CAMS monitors and forecasts European air quality and worldwide
long- range transport of pollutants.
3 | ELEMENT DEPOSITION IN GLOBAL
FORESTS
Various substances emitted from natural or anthropogenic sources
flow from the atmosphere into forest ecosystems by either wet or
dry deposition (Tørseth et al., 2012). Atmospheric deposition may
be harmful or beneficial for trees and other plants (Figure 2). Sulfur
(S) and nitrogen (N) compounds may function as either nutrients or
stressors for forests, even though they are derived from anthro-
pogenic air pollutants, such as sulfur oxides (SOx), nitrogen oxides
(NOx), and ammonia (NH3; Duan et al., 2016; Oksanen & Kontunen-
Soppela, 2021). When traveling through the canopy, acid deposi-
tion can cause direct damage to plant leaves (Du et al., 2017 ). When
deposited to the forest floor, N and S compounds are identified
as a cause of acidification and eutrophication (or N saturation) of
forest ecosystems (de Vries, 2021). Moreover, certain amounts of
phosphorus (P) and basic cations, such as calcium (Ca2+) and mag-
nesium (Mg2+), acting in forests as nutrients, are also derived from
anthropogenic emissions (Du et al., 2016, 2018). Climate change may
FIGURE 2 Main interactions of forest ecosystems with sulfur (S) and nitrogen (N) compounds. They may function as either stressors
(S) or nutrients (N), even when they are derived from anthropogenic air pollutants, such as sulfur oxides (SOx), nitrogen oxides (NOx), and
ammonia (NH3), with direct effects on forest canopy (Du et al., 2017 ) and indirect effects on acidification (Augustaitis et al., 2010) and
eutrophication (de Vries, 2021) including impacts on biodiversity (Clark et al., 2013), growth (Du et al., 2018), volatile emissions (Hansen
et al., 20 17; Liu & Greaver, 2009; Mushinski et al., 2 019; Schindler et al., 2020; Xie et al., 2018), and biogeochemistry (Gaudio et al., 2015;
Nakahara et al., 2010).
6 
|
    DE MARCO et al.
directly or indirectly affect the roles of these substances in forest
ecosystems (e.g., Mitchell & Likens, 2011; Nakahara et al., 2010).
Atmospheric deposition, especially of S and N compounds, has
declined over the last three decades (Sicard, De Marco, et al., 2016;
Tørseth et al., 2012; Zhong et al., 2020), despite many developing
nations still lacking effective SO2 emiss ion cont rol s. In Euro pe, depo-
sition of S and N peaked in the late 1970s and in the 1980s, respec-
tively (Engardt et al., 2017). In North America, deposition of S and
N peaked in the early 1970s and mid- 1990s, respectively (Mitchell
& Likens, 2011), when NH3 emission became more important
(Du, 2016). In South America, most average daily concentrations of
SO2 are below the World Health Organization air quality guidelines
(Peláez et al., 2020), and global atmospheric S deposition is lower
than in Europe, Asia, the United States, and Africa (Gao et al., 2018),
ranging around 4.96 ± 3.45 kgS ha−1 a−1 . In Asia, emissions of SO2
and NOx significantly increased from the early 1980s to the early
2000s (Ohara et al., 2007), 20 or 30 years later than in Europe and
the United States. The emissions of SO2 and NOx in China peaked in
2006 (Lu et al., 2011) and 2011– 2012 (Zheng et al., 2018), respec-
tively, and thereafter started decreasing. In China, emissions of NH3
reached a plateau in 1996 (Kang et al., 2016), although a gradual in-
crease in NH3 emissions in Asia (including China) was observed as of
2015 (Kurokawa & Ohara, 2020). A recent global analysis combined
inventory and modeling data to confirm that the total annual NOx
emissions finally stopped increasing in 2013, largely due to strict
control measures taken in China in recent years (Huang et al., 2017).
However, SO2 emissions in India overtook those in China in 2016 (Li
et al., 2017), and thus a focus should be placed on monitoring atmo-
spheric deposition in India and other developing countries. Major
air pollutants have been changing with industrialization in each re-
gion, from SO2 to NOX and NH3. With temporal changes of major
pollutants relative to industrialization, acidification, photochemical
formation of ozone, and excess N deposition appeared in sequence
as problems for forest ecosystems, as seen in the changes of main
causes of tree and forest decline in Northeast Asia (Takahashi
et al., 2020). Thus, emission reduction of S and/or N has been re-
flected gradually by the conditions of forest ecosystems.
In Europe and the United States, reduced S deposition resulted
in long- term declines in
SO2
4
concentrations in soil solutions (Berger
et al., 2016; Johnson et al., 2018) and freshwater (Garmo et al., 2014;
Vuorenmaa et al., 2017 ). However, since S compounds are retained
in forest ecosystems and released with changing environmental con-
ditions, changes in S leaching do not necessarily occur at the same
time as S deposition changes. Therefore, S output in forest catch-
ments often exceeds the atmospheric input due to legacy S pools
derived from past deposition (Vuorenmaa et al., 2017) or due to
changing climate (Mitchell & Likens, 2011), which might delay recov-
ery from acidification. In Asia, much of the deposited atmospheric
SO2
4
seems to be retained in forest soils (Duan et al., 2016; Sase
et al., 2019), which may imply a future risk of soil acidification under
changing climate. In fact,
SO2
4
concentrations and pH of river wa-
ters are related to the S emission/deposition rate (Duan et al., 2011;
Qiao et al., 2016; Sase et al., 2017, 2021). To understand the S cycle
in forest ecosystems, targeted studies on deposition trend and
changing climate are required (e.g., Mitchell & Likens, 2011; Sase
et al., 2019, 2021; Vuorenmaa et al., 2017 ).
Air pollution abatement may also reduce atmospheric inputs of
base cations (Tørseth et al., 2012), as reported for forest soil solu-
tions (Johnson et al., 2018) and freshwaters (Garmo et al., 2014;
Stoddard et al., 1999). Base cation nutrient s in China forest s neutral-
ized on average 76% of the potential acid load due to acid deposition
during 2001– 2015 (Du et al., 2018). Thus, base cation deposition
should be monitored simultaneously along with S and N deposition
as already done by several networks globally to assess nutrient sta-
tus and recovery from acidification in forest ecosystems.
Excess N inputs from the atmosphere have been disturbing
biogeochemical cycles in forest ecosystems (e.g., Aber et al., 1989;
Nakahara et al., 2010). With reduction in total N deposition mainly
due to NOX emissions, an improvement is expected in the NH.
However, high levels of NH3 deposition are still concerning because
NH3 emissions have not clearly reduced in many of the regions as
described above. Moreover, since emissions of SO2 and NOx have
been reduced resulting in significant decline of particulate forma-
tion (such as (NH4)2SO4 and NH4NO3), air concentrations of NH3
have been increasing and accordingly more localized NH3 deposi-
tion was identified in the United States (Butler et al., 2016). Even
though regional N deposition has gradually decreased, ecosystem
responses to N deposition appeared to show some degree of hyster-
esis (Gilliam et al., 2 019). In fact, there was no large- scale response
in understory vegetation, tree growth, or vitality to reduction of
N deposition in Europe, while a decline in
NO
3
concentrations in
soil solutions and foliar N concentrations were partly observed
(Schmitz et al., 2019). In Asia, three decades of increase in N depo-
sition in China have exerted significant impacts on soil and water
acidification, understory biodiversity, forest growth, and carbon se-
questration (Qiao et al., 2016; Tian et al., 2018). However, recovery
from acidification and N saturation has already started following a
reduction in N deposition in Japan (Sase et al., 2019), where high
S and N deposition and climatic anomalies caused acidification and
N saturation in the 1990s (Nakahara et al., 2 010). Nitrogen leach-
ing from forest ecosystems is controlled not only by N deposition,
but also by various factors, including tree age, forest management,
climate, and other limiting nutrients such as phosphorus. Moreover,
emissions of NH3 (e.g., Hansen et al., 2017), N2O (e.g., Schindler
et al., 2020; Xie et al., 2018), and NOy (as NO + NO2+ HONO; e.g.,
Mushinski et al., 2 019) as well as microbial nitrification rate (e.g.,
Fang et al., 2015) in forest areas should be taken into consideration
for actual N fluxes. Since N deposition may increase gas N emissions
from ecosystems (e.g., Xie et al., 2018), a comprehensive study con-
sidering bilateral N fluxes (both deposition and emission) should be
promoted to evaluate whether a forest ecosystem is a sink or source
of reactive N species.
The analysis of N dynamics in Latin America is complex, due to
the enormous diversity of unmanaged and managed ecosystems,
including arid deserts as well as temperate and tropical forests.
Cunha- Zeri and Ometto (2021) stated the major input of N in Latin
   
|
 7
DE MARCO et al.
American countries over the past decades occurred via natural bio-
logical fixation, compared to anthropic sources (fertilizers and fos-
sil fuel combustion). Nevertheless, human activities have currently
changed the N cycle of natural ecosystems in Latin America. For in-
stance, the conversion of unmanaged land to agriculture increased
biological N fixation up to twofold (Reis et al., 2020). Although the
highest total N deposition occurs in eastern and southern China,
Japan, Eastern US, and European forests, the highest dry deposi-
tion occurs in tropical forests (Schwede et al., 2018). For instance,
dry N deposition into the Atlantic Forest in the city of São Paulo
(Brazil) can exceed the critical N load found for most forests (Souza
et al., 2020).
Because of the continued increase in NH3 emission in some re-
gions (e.g., Kurokawa & Ohara, 2020) and stagnating values in oth-
ers (Maas & Grennfelt, 2016), N deposition is a per vasive issue that
impacts forest ecosystems. In addition, even relatively low levels of
N deposition affect the mycorrhizal association of trees (Lilleskov
et al., 2019; van der Linde et al., 2018) and may affect biodiversity of
sensitive species, such as lichens (Giordani et al., 2014). The magni-
tude and consequences of these human- induced changes in plant–
soil– microbe interactions as well as potential pathways for recovery
are currently open questions.
Moreover, excess N deposition may induce an imbalance of nutri-
ent ratios, such as N:P ratio (Krüger et al., 2020; Sardans et al., 2016).
However, the observational data on atmospheric P deposition are
still limited for forest areas (e.g., Chiwa, 2020; Du et al., 2016) and
N- P imbalances have been reported from various regions (Boccuzzi
et al., 2021; Krüger et al., 2020; Peñuelas et al., 2013). Taking into
account the global pattern of N and P limitation in forest areas (Du
et al., 2020), N and P deposition should be monitored together. Both
N and P cycles are listed as important Earth- system processes in the
concept of “Planetary boundaries” with N cycle already transgress-
ing its boundary (Rockström et al., 2009; Steffen et al., 2015).
Climate has an important role in regulating the global patterns of
terrestrial N and P limitation (Du et al., 2020). Specifically, there is a
shift from relative P to N limitation at lower mean annual tempera-
ture, temperature seasonality, mean annual precipitation, and higher
precipitation. Future climate change will likely reshape the spatial
pattern of nutrient limitation. For instance, climate warming will im-
prove N availability at mid- to- high latitudes via increasing biologi-
cal N fixation and N mineralization (Zaehle et al., 2010). Moreover,
growth stimulation by rising atmospheric CO2 concentration ([CO2])
will increase nutrient demand and, in turn, result in greater nutrient
limitation (Collalti et al., 2018; Wieder et al., 2015). The changing nu-
trient status under climate change will likely interact with the effects
of S and N deposit ion and thu s they should be considere d simulta ne-
ously when projecting future forest dynamics.
4 | GROUND- LEVEL OZONE IMPACTS
Background O3 concentrations have increased throughout the last
century due to the rising anthropogenic emissions of O3 precursors
from fossil fuel and biomass burning (Cooper et al., 2014; Monks
et al., 2015), although volatile organic compounds (VOCs) also are
major precursors (Wei et al., 20 14). Despite the decreasing trend of
other air pollutants in the last decades (e.g., S and N compounds,
heavy metals), global- scale background O3 concentrations increased
(Jakovljevet al., 2021; Sicard, 2021), but slight regional- scale de-
creases in peak concentrations were observed (Schaub et al., 2018).
Thus, O3 is nowadays one of the main phytotoxic air pollutants with
the potential to affect forest ecosystems worldwide (Agathokleous
et al., 2020; Bytnerowicz et al., 2016; De Marco et al., 2020; Feng,
Shang, Gao, et al., 2 019; Sicard, Augustaitis, et al., 2016).
Ozone burdens are higher in the Northern (O3 mean concentra-
tion 35– 50 ppb) than in the SH (O3 mean concentration <20 ppb;
Sicard et al., 2 017). For example, widespread O3- induced visible in-
jury, a specif ic damage ass ociated wit h O3 exposure, was foun d at 17
forest plots in Europe (Paoletti et al., 2019; Sicard et al., 2020). The
NH is more covered by land and terrestrial ecosystems, and more
inhabited by humans than the SH, and thus is more affected by an-
thropogenic activities. However, the SH is less monitored and thus
O3 burdens and effects may be underestimated. While there are
hundreds of papers on O3 effects on forest plants and forests in the
NH (i.e., Agathokleous et al., 2015; Feng, Shang, Gao, et al., 2019;
Izuta, 2017; Sicard et al., 2020) indicating various effects of O3 in
interaction with climate change (Figure 3), relevant research in the
SH remains scarce.
The analysis of O3 effects in Latin America is complex due to
the enormous diversity of natural and agricultural ecosystems.
Most monitoring studies on O3 effects on forest plants conducted
in the SH come from Brazil. Urban and industrial development has
been more intense along the Atlantic Brazilian coast, especially in
Southeastern region. Consequently, more severe O3 effects on the
Atlantic forest located in this subtropical region (mainly São Paulo
and Rio de Janeiro States) are expected (Domingos et al., 2003;
Moura et al., 2014, 2018). Ozone effects on native tree species from
the Atlantic Forest have recently been determined in the field or
experimentally, pointing to distinct tolerance levels and highlighting
the need to expand knowledge on this topic (Cassimiro et al., 2016;
Engela et al., 2021; Fernandes et al., 2019; Moura et al., 2018). In
the SH, the Amazon spans over 629 million hectares of rainfor-
est, accounting for 54% of the total rainforests left on Earth (Peng
et al., 2020). Recent modeling approaches have shown O3 concentra-
tions have increased above the Amazon and Cerrado biomes in Brazil
as a response to biomass burning and regional air pollution (Gerken
et al., 2016; Pope et al., 2 019). The lowest O3 exposures reported
are in Australia, New Zealand, southern parts of South America, and
some northern parts of Europe, Canada, and the United States (Mills
et al., 2018; Sicard et al., 2017). However, unfortunately, a proper O3
monitoring network does not currently exist. Despite the presence
of ground- level O3 monitoring networks in all the developed coun-
tries (Lefohn et al., 2018), there is still a lack of an integral network
of ground- level O3 monitoring across Asia, although 1500 monitor-
ing stations have recently been installed in China (Feng, Shang, Gao,
et al., 2019).
8 
|
    DE MARCO et al.
Another challenge in monitoring O3 impacts on forests is the
choice of metrics. The AOT40 index (Accumulated Ozone over
Threshold of 40 ppb ozone), describing the exposure of plants to high
O3 concentrations, is the default measure for policy direc tives of the
European Union (Directive 2008/50/EC). However, AOT40 has been
criticized because it is not a proxy of gas uptake through leaf sto-
mata (stomatal flux), and flux- based indices have been applied (Anav
et al., 2022; De Marco & Sicard, 2 019; Paoletti et al., 2019; Sicard
et al., 2020) and showed O3 risks to vegetation would be different
from AOT40 (Anav et al., 2016; De Marco et al., 2015). The new
standard developed in Europe (Emberson et al., 2000) is the stoma-
tal O3 flux, defined as POD (Phytotoxic Ozone Dose). This standard
depends not only on O3 concentration, but also environmental (e.g.,
light intensity, air temperature, relative humidity, soil moisture) and
plant conditions (phenology, leaf morphological, and physiological
traits). A major impact of O3 is reduced aboveground and below-
ground carbon sequestration of forests (Agathokleous et al., 2016;
Gao et al., 2017; Figure 2). Ozone effects on biogenic volatile organic
compounds (BVOCs) are complex, as some compounds may decrease
(e.g., isoprene) while other compounds increase (e.g., monoter-
penes; Feng, Yuan, et al., 2019). Different BVOC compounds have
different capacity to generate O3, with isoprene having higher O3-
forming potential than monoterpenes (9.1 g 03 (g VOC)−1 and 3.8 g 03
(g VOC)−1, respectively; Benjamin & Winer, 1998). However, sesqui-
terpenes and some monoterpenes also contribute to the removal of
O3 at the canopy level and play an important role in the feedback
between stress- induced VOC emissions and O3 or aerosol forma-
tion (Calfapietra et al., 2013). The emission of isoprene, the most
abundant BVOC, can also be decreased by drought and CO2 and in-
creased by warming (Feng, Shang, Li, et al., 2019), indicating complex
O3- climate interactions that remain elusive in real- world forests. Soil
microbial processes contribute to emission of BVOCs and NOx that
act as O3 precursors (Gray et al., 2010). Overall, soils play an im-
portant role in forest VOC exchange, defining also carbon storage
by forest ecosystems, and fluxes depend upon BVOC compounds
and vegetation types (Mäki et al., 2019; Rinnan & Albers, 2020; for
details and values of fluxes in different vegetation types and envi-
ronmental media, see also Tani and Mochizuki (2021)). However, the
FIGURE 3 The Gordian Knot of the Forest– Ozone– Carbon interactions. In the pre- industrial epoch, carbon is stored via photosynthesis
(1) and leads to long- term carbon sequestration into aboveground and belowground (roots and soil) wood biomass (2) (Agathokleous
et al., 2016; Grantz et al., 2006). The higher CO2 levels, alone, in the atmosphere are expected to “feed” forest growth (Koike et al., 2018)
and have beneficial effects. The increased O3 levels, alone, depress forest trees, contributing to “forest decline syndrome,” that is, visible
injury, photosynthesis, carbon sequestration, carbon storage changes (7), and biomass decay, which also releases CO2 in the atmosphere (8)
(Agathokleous et al., 2016; Sandermann et al., 19 97; Sicard et al., 2021; Takahashi et al., 2020). In a positive feedback, the depressed forest
vegetation emits more BVOCs (4), further increasing O3 levels (Peñuelas & Staudt, 2010). Concurrent elevated concentrations of CO2 end
O3 may outcome to a sustained increase in Net Primary Productivity (NPP), while the adverse long- term effect of increased O3 on NPP may
be lesser than projected (Talhelm et al., 2 014). Elevated CO2 levels negate or even overcompensate the negative O3 effect on ecosystem
functions and the cycles of carbon and nitrogen. Anthropogenic emissions of CO2, NOx, and volatile organic compounds (VOCs) (3) as well
as biogenic VOCs (BVOCs) emitted by forests (4) contribute to increased O3 levels in the atmosphere (Yu & Blande, 2021). Soil microbial
processes contribute to soil- emitted BVOCs and NOx (O3 precursors; Gray et al., 2010 ) as well as CO2, N2O and CH4 (Yao et al., 2009; Zhang
et al., 2021) (5). Under advanced climate change, forest fires are expected to be more frequent and larger than in the pre- industrial epoch
(Zhang et al., 2021). These fires release carbon monoxide (CO), organic carbon (OC), NOx (all of which contribute to O3 formation), and
black carbon (BC; which influences photosynthesis by increasing diffuse radiation) as well as CO2 (which further intensifies global warming;
Flannigan et al., 2009; Kumar et al., 2019 ; Pellegrini et al., 2021; Yue & Unger, 2018) (9).
   
|
 9
DE MARCO et al.
specific contribution of soil in VOC exchanges and O3 formation re-
mains poorly understood.
5 | TRACE ELEMENTS AND R ADIOACTIVE
CONTAMINATION OF FOREST ECOSYSTEMS
Heavy metal pollution was an important subject in widespread forest
decline in the 1980s– 1990s (Gawel et al., 1996), but more recently has
become a major item in phytoremediation (Pulford & Watson, 2003)
and environmental monitoring (Godzik, 2020). The term “heavy met-
als” is now discouraged, and these elements are now included more
broadly as “trace elements” (Pourret & Bollinger, 2018). Trace ele-
ments are a major component of particulate pollution (Antoniadis
et al., 2017; Grantz et al., 2003; Li et al., 2015; Schlutow et al., 2021;
Tóth et al., 2016). At the global scale, trees are important for their
role in retaining particulates (Yue et al., 2021). Nevertheless, in
some regions, soil contamination by trace elements remains so high
that it continues to kill trees and prevents natural recovery (Kozlov
et al., 2009). Among trace elements, radionuclides display the most
phytotoxic potential.
The use of nuclear energy or nuclear applications in health, ag-
riculture, environmental management, or industry/military resulted
in releases of radionuclides into the environment (Hong et al., 2012).
The first large- scale radioactive contamination from anthropogenic
sources occurred through global radioactive fallout from nuclear
weapons' tests conducted in the atmosphere during 1945– 1980
(Aoyama et al., 2006; United Nations, 2000). A variety of long- and
short- lived radionuclides were released during nuclear incidents; in
particular 137Cs with a relatively long half- life (~30 years) compared
to other radionuclides, such as 134Cs and 131Cs. Other major releases
of radionuclides occurred from the Chernobyl nuclear power plant
accident in 1986 (International Atomic Energy Agency, 2006) and
from the Fukushima Daiichi Nuclear Power plant accident in 2011
(Chino et al., 2011; Terada et al., 2020; Yoshida & Takahashi, 2012).
Radioactive contamination of forests has different types of
impacts (Figure 4). First, direct radiation can affect trees and an-
imals and occur at the level of DNA, cells, individuals, population
to whole ecosystems, and ranges from reparable DNA damage to
death of organisms (Committee on the Biological Effects of Ionizing
Radiation, 1990 ). An example of direct impacts of high radiation
doses to trees is the “Red forest” in the Chernobyl exclusion zone,
where pine trees became reddish brown and died following the ac-
cident (Beresford et al., 2016). Another visible impact of radiation
exposure in trees is the occurrence of morphological abnormalities
(Watanabe et al., 2015; Yoschenko et al., 2011, 2016). Compared
to the effects caused by high doses of radiation, those poten-
tially caused by relatively lower radiation dose are confounded by
many other factors and are still not clearly understood (Beresford
et al., 2020; Ji et al., 2019; Strand et al., 2017). In exposed areas, for-
est ecosystems are released from pressure by human existence, re-
sulting in creation of ecological niches and expansion of populations
of some species (Deryabina et al., 2015; Lyons et al., 2020; Perino
et al., 2019). Through intensive monitoring, it was confirmed that the
overall dynamics of 137Cs within forest ecosystems were similar be-
tween Chernobyl and Fukushima: tree canopies captured the depo-
sition of 137Cs and 137Cs migrated from the canopy to the soil surface
via water and litter fall, and most of it stays in the top layers of soil
(Itoh et al., 2015; Kato et al., 2019; Suchara et al., 2 016). However,
the migration velocity and distribution patterns of 137Cs within for-
ests and tree bodies differ substantially among forests and trees
(Imamura et al., 2017; Ohashi et al., 2017). It is essential to continue
experimental studies to identify the key processes influencing 137Cs
dynamics in forest systems, such as soil potassium concentrations
and fixation processes within soils (Kobayashi et al., 2019 ; Manaka
et al., 2019). Various models have been developed to characterize
137Cs dynamics in forests; however, improvements are necessary
to reproduce variations between forest types and species compo-
sitions (Hashimoto et al., 2020). Another aspect of radionuclide pol-
lution is that deposited radionuclides, which are easy to detect and
measure, provide an unintentional but useful opportunity to track
biogeochemical cycles in forest ecosystems (Fukuyama et al., 2008).
6 | COMBINED EFFECTS OF MULTIPLE
FACTORS ON FOREST ECOSYSTEMS
Our knowledge on combined effects of multiple factors on ecosys-
tem health originated primarily from temperate and boreal forests
of North America and Europe and is limited for tropical forests,
FIGURE 4 Diagram of direct and indirect effects of forest
radioactive contamination. The deposited radionuclides remain in
the forest and continue to circulate in the forest ecosystem, and
radiation can have adverse effects on forest biota (direct effects).
Restrictions on forest use and land abandonment to avoid exposure
can also affect forest ecosystems, including changes in vegetation
and wildlife populations. It has direct and indirect impacts on
ecosystems and local residents.
10 
|
    DE MARCO et al.
especially of those in Africa (Matyssek et al., 2 017). In other words,
areas that have recently experienced the highest risk of forest deg-
radation are studied to a lesser extent than the areas where risk is
low. In addition, many communities whose food security and wealth
generation critically depend on forests are located in geographic re-
gions where our understanding of factors affecting forest ecosys-
tem health is poor. This geographic bias is typical for ecological and
environmental research (Archer et al., 2014), and its consequences
are generally seen as severe, because results obtained with one
study system may appear of little use in predicting the responses of
another, geographically distinct, study system (Haukioja et al., 1994).
Air pollution levels may become more harmful for plants as the
climate warms (Zvereva et al., 2008, 2010). More multi- factorial
manipulative studies are needed because effects of two or more
co- occurring factors on tree growth and forest productivity can-
not be adequately predicted from single- factor experiments
(Niinemets, 2010). The combined effects of two major abiotic aspects
of global change, mostly changes in CO2 and warming, on growth of
forests are studied in detail (Baig et al., 2015; Curtis & Wang, 1998;
Zvereva & Kozlov, 2006), and suggest air temperature may modify
plan t responses to elevate d CO2. Across 42 experiments with woody
plants, aboveground biomass increased significantly with both CO2
(the so called “fertilization effect”) and air temperature (by 21.4%
and 18.1%, respectively), whereas these two factors acting simulta-
neously showed a much smaller effect (8.2%) because of compen-
sating effects (Baig et al., 2015). Nitrogen fertilization enhances the
biomass response to elevated CO2 (Parrent & Vilgalys, 2007) despite
not universally (Terrer et al., 2019). The type of mycorrhiza was also
an important factor related to the effects of soil nutrient availability
on elevated CO2- induced growth enhancement (Baig et al., 2015).
However, two- factorial experiments involving both O3 exposure
and elevated CO2 are limited. Several studies under elevated CO2
showed a reduction in the negative effects of O3 because elevated
CO2 induced stomatal closure leading to lower O
3 uptake (Grams
et al., 1999; Watanabe et al., 2017 ). In contrast, the addition of N
alone exacerbated negative effec ts of O3 on photosynthesis of trees
(Feng, Shang, Li, et al., 20 19), while exposure to drought stress did
(Gao et al., 2 017) or did not protect plants from O3- induced effects
(Alonso et al., 2003, 2014).
Forest health also can be compromised by insect herbivory, in-
cluding both devastating outbreaks of forest pests and changes in
background herbivory. Despite relatively low levels of plant damage
(5%– 7% of leaf biomass annually: Kozlov et al., 2015), background
herbivory greatly reduces growth of woody plants (Shestakov
et al., 2020; Zvereva et al., 2012). Although warming, drought, CO2
increases, N deposition, and air pollution were repeatedly found to
increase herbivory (Lincoln et al., 199 3; Logan et al., 20 03), these
conclusions were likely affected by research and publication biases
(Zvereva & Kozlov, 2010) and/or were derived from results of short-
term laboratory experiments, which tend to overestimate the ef-
fects relative to natural ecosystems (Bebber, 2021). Within forest
ecosystems across the globe, no increase in insect herbivory was
observed from 1952 to 2013 (Kozlov & Zvereva, 2015). Similarly,
long- term monitoring did not reveal the effects of either pollution-
induced disturbance or 2.5°C climate warming on insect herbivory
in subarctic birch forests (Kozlov et al., 2017). Thus, the evidence
regarding combined effects of climate warming and air pollution on
insect herbivory remains somehow contradictory.
Other factors whose effects on forest trees have been stud-
ied in multi- factorial studies include (but are not limited to) cattle/
deer grazing, harvest of non- timber forest products, drought, flood-
ing, soil salinization, spring frost, heat waves, and increased ultra-
violet radiation (e.g., Mac Nally et al., 2011; Pliūra et al., 2019; Sugai
et al., 2019; Varghese et al., 2015). However, a low number of such
studies precludes any generalization regarding effects of these fac-
tors, combined with CO2 and air temperature increases or O3 and
insect herbivory on health of forest ecosystems. Modeling studies
jointly assessing the effects of climate change and air pollution can
greatly help for understanding and predicting future developments
of forests (Akselsson et al., 2016; Dirnböck et al., 2017; Etzold
et al., 2020; Fleck et al., 2 017; Rizzetto et al., 2016).
7 | GENETIC INFORMATION RELATED
TO PHENOTYPES AND PHYSIOLOGY OF
FOREST TREES
Air pollution, climate change, increased pests and pathogens, land-
use changes, and forest fragmentation can all reduce genetic diver-
sity and make forests more fragile and sensitive to other threats
(Gauthier et al., 2015). Current vegetation and forest growth models
are largely parameterized on direct growth and gas exchange meas-
urements or remote sensing, while information from biological and
genetic regulation mechanisms are still scarce. For example, part of
the carbon fixation products (i.e., photosynthates) that is not used
for biomass production is released in soil as root exudates, some is
stored, and some organic carbon is emitted as BVOCs affecting plant
and community ecology and atmospheric chemistry (Blande, 2021;
Collalti et al., 2020; Maja et al., 2015; Naidoo et al., 2019; Šimpraga
et al., 2019). Carbon sink strength of trees is known to be impaired
by limitations in water and nutrient availability, heath spells, air
pollutants, and increased herbivory. However, plant defense pro-
cesses against different abiotic and biotic factors are complex and
involve multiple signaling pathways (He et al., 2018), potentially af-
fecting how carbon is allocated to different organs (Merganičová
et al., 2019). Most of the underlying resistance mechanisms are de-
scribed or predicted from short- living herbaceous model systems,
whereas investigations on mechanisms of defense and adaptation
of forest trees are much more challenging due to long lifetime, high
genetic diversity, and variation of growth environments and climates
(Naidoo et al., 2019). There is an urgent need to intensify studies on
the mechanisms underlying the resilience of forest ecosystems to
current and long- term effects of air pollution and climate change,
utilizing genetic, species, and ecosystem- level functional diversity
as well as adaptive management, resistance breeding, and genetic
engineering (Naidoo et al., 2 019). Mechanistic understanding is
   
|
11
DE MARCO et al.
increasingly important also for efforts in afforestation and protec-
tion of primar y forests. In principle, there are two main approaches
for achieving resistance in forest trees: (i) selection of resistant
phenotypes identified in field experiments (Sniezko & Koch, 2017)
or polluted sites (Eränen et al., 2009; Kozlov, 2005); and (ii) struc-
tured breeding programs relying on multitude of omic techniques
(Naidoo et al., 2019). The databases for genetic information of tree
species have been rapidly increasing, and the most important model
systems for forest trees are Populus, Eucalyptus, Quercus, Castanea,
Pseudotsuga, Pinus, Picea, and Betula genuses (Falk et al., 2018;
Salojärvi et al., 2017). Genetic engineering efforts by forest bio-
technology companies have produced transgenic Eucalyptus and
Populus trees with enhanced growth and disease- resistant proper-
ties (Naidoo et al., 2019). Silver birch (Betula pendula Roth) is an ex-
cellent model system for elucidating the adaptation and acclimation
capacity of forest trees to rapidly changing climate due to its (i) wide
latitudinal and longitudinal distribution; (ii) recent advances in popu-
lation genomics and evolutionary history of birch species (Salojärvi
et al., 2017); and (iii) existence of well- characterized birch genotypes
that have been intensively studied for C and N economy, photosyn-
thetic efficiency, metabolism, chemistry, and phenology (Deepak
et al., 2018; Tenkanen et al., 2020). The population genomic analyses
of silver birch provide insights on natural selection mechanisms, with
candidate genes relevant for adaptation of trees to changing envi-
ronment, biotic stress, and growth regulation (Salojärvi et al., 2017 ).
Studies with birch have also shown the C- sink strength of trees can-
not be explained by physiological or genetic approaches alone, but
there are many negative and positive interactions with pollutants,
climate, pests, pathogens, microbiomes, and between plants that
should be understood in more detail (Naidoo et al., 2019; Silfver
et al., 2020; Wenig et al., 20 19).
Plant phenotypes are strongly affected by the environment, and
often genotype per environment interaction is the factor of greatest
interest. Methodologies have been developed for non- destructive
forest- level and individual tree- level phenotyping with remote sens-
ing techniques, which are particularly useful for identifying superior
genotypes under different stress conditions (Dungey et al., 2018;
Kefauver et al., 2012; Ludovisi et al., 2017 ). Recent advances in
metagenomics and the increasing knowledge of the importance
of microbiomes in plant health offer new opportunities for forest
health management (Imperato et al., 2019; Naidoo et al., 20 19;
Wenig et al., 2019). The regulatory networks of forest trees and the
beneficial non- pathogenic microbes living around and on the sur-
faces of plant roots (rhizosphere), leaves (phyllosphere), or in the
internal plant tissues (endosphere) can be particularly important for
carbon and nutrient dynamics of trees and the development of tree
immunity (Naidoo et al., 2019). Microbes are known to help plants
in water and nutrition acquisition, defense against pathogenic mi-
crobes, tolerance to abiotic stress, adaptation, promotion of the es-
tablishment of mycorrhizal association, and plant growth regulation,
forming a holobiont system with host trees (Imperato et al., 2019;
Naidoo et al., 2019; Wenig et al., 2019). Fungal and bacterial com-
munities in forest soils have been shown to respond to changes in
climate with a shift in their community composition as well as in their
diversity (Dubey et al., 2019; Jansson & Hofmockel, 2020; Milović
et al., 2021; Simard, 2010). For example, under elevated CO2, we
can observe alteration in relative abundances of bacteria and in-
creased bacterial to fungal ratio (Dubey et al., 2019), as well as an
increase in ectomycorrhizal colonization rate but a decrease in ecto-
mycorrhizal diversity (Wang et al., 2015). Warming and elevated O3
reduced ecto- and arbuscular mycorrhizal colonization and shifted
arbuscular mycorrhizal community composition in favor of the genus
Paraglomus, which has high nutrient- absorbing hyphal surface (Qiu
et al., 2021; Wang et al., 2015). At the same time, exposure to higher
levels of O3 is associated with lower soil microbial biomass and with
changes in the overall structure and composition of poplar rhizo-
sphere soil microbial communities (Li et al., 2021). The decreased
growth of roots and decrease in ectomycorrhizal colonization rate
and a shift in species abundance might be an early indicator of the
damaging impacts of O3 in some tree species, occurring prior to visi-
ble responses of aboveground tree parts (Katanić et al., 2014).
8 | MODELING FOREST ECOSYSTEMS FOR
RISK ASSESSMENT
Scientific methods in forestry, including empirical models of tree
growth, were primarily used for optimization of timber harvest
throughout the 20th century (Por& Bartelink, 2002). A more in-
tegrative modeling approach, acknowledging natural disturbances
(e.g., wind, fires, pests, diseases) as inherent elements of forest
ecosystem dynamics, was developed when computational advance-
ments allowed for the integration of greater complexity (Blanco
et al., 2020; Perera et al., 2015), although some biotic factors of for-
est disturbance such as herbivory are still rarely modeled (De Jager
et al., 2017). Since the forest dieback and acidification debate in
Europe in the 1980s, large efforts were put into improving under-
standing and prediction of anthropogenic disturbances on biogeo-
chemical dynamics of forests. Starting from models mainly targeting
the fate and effects of acid rain in forest ecosystems (Nilsson, 1988;
Sverdrup & De Vries, 1994), simulation tools have broadened to in-
clude other pressures, such as N deposition (De Vries et al., 2010),
O3 (Hoshika et al., 2015), and climate change and forest management
(Collalti et al., 2018). Modeling has been demonstrated to be a valu-
able tool for studying forest responses to present and future distur-
bances, allowing ecologists and foresters to deal with the study of
complex interactions and to evaluate future management strategies
(e.g., Collalti et al., 2018; Fleck et al., 2017) or policy options (e.g.,
Belyazid et al., 2010; Dirnböck et al., 2018).
Existing relationships between forest structure and compo-
sition and environmental variables were initially used to build
empirical models that describe past ecosystem behavior and ex-
trapolate to future conditions (Gustafson, 2013). Subsequent
modeling efforts simulated the causal biogeochemical mecha-
nisms that underlie the responses of ecosystems to these envi-
ronments (Kimmins et al., 2008). These so- called process- based
12 
|
    DE MARCO et al.
models (PBMs) study the ecological processes and are considered
one of the most reliable approaches for modeling forest eco-
system dynamics under global change (Evans, 2012; Maréchaux
et al., 2020). However, forecasting forest growth is still a priority
in many studies, either for planning forestry activities under air
quality and climate change scenarios or as part of carbon storage
calculations (Blanco et al., 2020).
The general trend toward biodiversity conservation in interna-
tional policies (e.g., EU Biodiversity Strategy for 2030; European
Commission, 2020), its importance in preserving ecosystem ser-
vices, and the use of biodiversity metrics as indicators in risk
assessment (Coordination Centre for Effects, 2017) and pol-
icy evaluation (Hein et al., 2018) make the simulation of species
composition changes a decisive function for any model. When
dynamic PBMs are used for forecasting biodiversity shifts, they
are usually combined with vegetation response models based on
species niche suitability and competition (Belyazid et al., 2019;
Dirnböck et al., 2018). PBMs have been used at stand (e.g., Collalti
et al., 2016), landscape (Shifley et al., 2017), regional (Belyazid
et al., 2 019; De Marco et al., 2020; Santini et al., 2014), and global
scales (e.g., Krause et al., 2017 ). However, their implementation is
restricted at larger scales since PBMs need large, detailed input
datasets, which are often not available at national or continen-
tal scales. At these scales, a currently suitable approach is using
new models, mostly empirical, based on currently available large
datasets, such as species distribution models (SDMs; Maréchaux
et al., 2020; Noce et al., 2017). In the same way that PBMs rose
with the increasing computational power during the last decades
of the 20th century, SDMs have improved during the present de-
cade, in parallel to the increase in web available, reliable spatial-
referenced data, including environmental and meteorological
data, forest inventories, habitat distribution, aerial images, and
remote sensing (Pecchi et al., 2019; Urban, 2015). SDMs are usu-
ally statistical models that are currently used to support sustain-
able planning of forests at national and international scales (Zang
et al., 2012), with correlative SDMs using maximum entropy al-
gorithms being most frequently used (Noce et al., 2017; Pecchi
et al., 2019). Using forest decision support systems, climate
change scenarios and the balance of delivered ecosystem services
can be suggested as a methodological framework for validating
forest management alternatives aiming for more adaptiveness in
sustainable forestry (Marano et al., 2019; Mozgeris et al., 2019).
Mo reov er, some of the ve get ati on model s assoc iated wit h PBMs to
assess or forecast biodiversity are SDMs that may be applied from
site to regional scales (e.g., Wamelink et al., 2020). There are some
recent examples of SDMs implemented to assess forest biodiver-
sity respons e to at mospheri c pollution and climate change, such as
Hellegers et al. (2020) and Wamelink et al. (2020). However, these
models still lack essential information to feed their predictions,
since new field observations and experiments with novel set- ups
(e.g., Hansen & Turner, 2019) are needed to address the potential
successional and disturbance dynamics under the forthcoming cli-
mate conditions (McDermott, 2020). Therefore, there are several
possible approaches for different problems that scientists and
managers must deal with (Blanco et al., 2020; Fabrika et al., 2 019;
Maréchaux et al., 2020). The modeling process might be as com-
plex as needed by risk assessment objectives (Figure 5), providing
models and data are available and suitable. In general terms, empir-
ical models are good at predicting biomass and forest structure in
the shorter term, and consequently producing good management
recommendations for the present conditions, but are not reliable
in novel situations (i.e., future air pollution and climate change).
PBMs are good at studying effects and underlying processes of
change, particularly in the context of global change. However,
they still have low feasibility at broad scales, and the calibration
and validation processes are highly time- consuming (particularly
for the less- modeled species or regions). SDMs are appropriate for
early risk assessment on biodiversity conservation at the broad-
est scales, but still too empirical, which diminishes their reliability
in the long term (Urban et al., 2016). Mixing process- based with
empirical approaches (hybrid models), integrating, and connecting
different models (meta- and mega- models; Blanco, 2013) are ex-
cellent strategies to answer specific questions.
9 | NEW DIRECTIONS: INTEGRATING
EXPERTS' OPINIONS
9.1 | Air pollution monitoring network
While conventional field- based monitoring plots will continue
to dominate the mainstay of air pollution and climate change re-
search in forests, they are costly and often logistically difficult to
conduct over large areas. Therefore, remote sensing techniques
will be more and more appropriate for large- scale monitoring pro-
grams, even though a more in- depth approach still needs to be
developed. Finer temporal intervals are required for in- depth un-
derstanding of some responses (e.g., stomatal O3 fluxes require
a continuous- monitoring approach). Highly instrumented field
sites are now cheaper and technically affordable. Integrating data
from transcontinental long- term ecological research infrastruc-
tures in tree- based models would lead to a better understanding
of how ecosystems work (Fischer et al., 2011). Long- term data
series can be integrated in existing big databases such as the
Global Atmosphere Watch (GAW) Program and the international
Tropospheric O3 Assessment Report (TOAR; Schultz et al., 2017;
WMO, GAW, 2003). These raw databases can lead to the devel-
opment of new products for temporal and spatial analysis (data
analysis, maps of data distributions, and data summaries) that are
freely accessible to the scientific community and other stakehold-
ers. Such databases can be used as tools for mechanistic and diag-
nostic understanding and upscaling.
The need for a global forest monitoring is irrefutable, and su-
persites” promote the integration of research communities in a
transcontinental collaboration network by upgrading existing
ground- based observation networks (e.g., FLUXNET, ICP, NEON)
   
|
13
DE MARCO et al.
covering all biogeographic areas (e.g., tropics, subtropics) and eco-
system types (e.g., woody savannas).
9.2 | Elements deposition in forests
The effects of S and N deposition on forest health have been
reducing gradually in many regions but problems have not been
solved. Legacy S pools remain, which could be affected by chang-
ing climate. Reduction of S deposition is associated with reduc-
tion of base cation deposition, which may alter nutrient status
and increase the risk of further soil acidification. The total inor-
ganic N deposition has been declining due to the implementation
of air pollution control policies, but the relative importance of
NH3 emissions and deposition is now higher (Butler et al., 2016;
Du, 2016), showing a relative increase of 0.38% per year over the
period 1985– 1999 (Du, 2016). Since ecosystem responses to de-
clining N deposition may show hysteresis (Gilliam et al., 20 19) and
key mechanisms of the N- induced changes in forest ecosystems
are not fully understood (Lilleskov et al., 20 19), long- term monitor-
ing of N- , S- , and P- cycles and base cations deposition should be
studied together to better understand biogeochemical processes
and plant biodiversity under climate change. Moreover, interac-
tions between nutrient deposition and rising O3 concentrations
should be considered in future studies (Shi et al., 2017). Long- term
monitoring should be continued even after significant air pollution
reductions to capture and understand the potential long- term ef-
fects of pollution and ecosystem recovery.
9.3  | Ground- level ozone
Surface O3 concentrations are generally higher in rural areas than in
urban areas (Sicard, 2021). However, as O3 levels are rising in cities
(Sicard, 2021), special attention should be paid to urban and peri-
urban forests, which offer services to local communities (Bruckman
et al., 2016) and can help meet air quality standards in cities (Sicard
et al., 2018). Because forest tree species play important (species-
dependent) dual roles as sinks and sources of O3 precursors (Geng
et al., 20 11; Saitanis, Agathokleous, et al., 2020), the O3 forming
potential (OFP) of the best regionally adapted forest tree species
should be investigated and taken into account by decision- makers to
select species with lower OFP for urban planning (Sicard et al., 2018).
The observed high O3 burdens, their high spatial heteroge-
neity, and the differential susceptibility of forest tree species to
O3, as well as their dual role as O3 sinks and precursor sources
(Agathokleous et al., 2020; Li et al., 2018), suggest an urgent
need for the establishment of a globally denser O3 monitoring
network in natural forest ecosystems in particular in the SH.
A new approach to the global O3 monitoring network and alter-
native methods for monitoring O3 are feasible thanks to innova-
tive technologies (Saitanis, Sicard, et al., 2020), which will help
FIGURE 5 Simplified flux of information diagram in modeling approach for risk assessment of air pollution and climate change. The
modeling and risk assessment process might be as complex as the modelers need and the availability and suitability of models and data allow.
There are three main blocks (grey bands): atmospheric and ecological models (top two) are the tools to reach the objective of risk assessment
(bottom). Models are used both in the internal processing and description of the information contained in the boxes, and in the transmission
of information between them (model inputs and outputs). For example, habitat suitability can be modeled using vegetation response models
(such as VEG; Belyazid et al., 2019 or PROPS; Dirnböck et al., 2018), that are particularly designed to process output from process- based
models as input information, but it can be also modelled by species distribution models that are particularly designed for large- scaled input
datasets (e.g., Noce et al., 2017). Information generally needed in any environmental study (such as soil and terrain variables) has been
obviated in this diagram.
14 
|
    DE MARCO et al.
to understand combined effects of O3 with other emerging envi-
ronmental factors. There is also an urgent need to generate flux-
base d standards and criti cal levels for fo res t prote ction taking into
account the sensitivity of dominant forest tree species. Because of
its limitations, the AOT40 index should not be adopted as default
for risk assessment (Agathokleous et al., 2019; Anav et al., 2022;
Sicard, Augustaitis, et al., 2016). Finally, the development of
countermeasures for controlling anthropogenic O3 precursor
emissions is also urgently needed.
Further research is still needed to develop O3- effect indicators
related to other ecosystem ser vices provided by forests such as bio-
diversity, soil protection, and water conservation. Nonlinear models
should be used for establishing cause– effect relationships under
experimental conditions (e.g., Agathokleous et al., 2019 ; De Marco
et al., 2013).
9.4 | Multiple stressors on forest ecosystems
For a better knowledge on combined effects of multiple factors on
ecosystem health, the selection of tree species for future studies
should account for their phylogenetic relatedness with already stud-
ied species. Ecological and environmental studies addressing the re-
sponses of tropical forests to combined effects of climate change and
air pollution should be intensified, in particular in areas at higher risk
of deforestation in the SH. This research domain is strongly biased
toward temperate and boreal forests of the NH. The evolutionary
changes in response to rising global CO2 levels and air temperature
elevation are known to occur in some plants, but the contribution of
evolutionary processes to the forest responses to steady CO2 and
air temperature rises remains unexplored. Experimental studies, ad-
dressing combined effects of different abiotic factors on forests,
should be intensified in the SH and should carefully select tree spe-
cies to assure a better representation of taxonomic and functional
diversity of the approximate 73,000 tree species now found on the
Earth (Cazzolla Gatti et al., 2022).
9.5 | Radioactive contamination of
forest ecosystems
Despite many papers reporting radioactivity effects on forest eco-
systems (Strand et al., 20 17; Tamaoki, 2016), there is still no con-
sensus on the mechanism through which radiation impacts forest
ecosystems or the dose rates at which impacts begin to occur
(Beresford et al., 2020; Strand et al., 2017). More robust and syn-
thesis studies are essential to inform (i) key processes regulating
the dynamics of radionuclides within forests; (ii) models for track-
ing radionuclides and prediction; (iii) holistic assessment of impacts
caused by radioactive contamination and its countermeasure devel-
opment; and (iv) use of 137Cs as a tracer. Furthermore, cost efficient
forest countermeasures must be developed and decisions must in-
clude locals, scientists, stakeholders, and governments.
9.6  | Genetic information of forest trees
More effort should focus on phenomics, combining high- throughput
capture of tree phenotypes, genotype information, data science,
and engineering (Falk et al., 2018; Naidoo et al., 2019). Future work
should include metadata integration and improved visualization for
comparative genomics. Characterizing the root traits and pheno-
types with association to genomics and shoot phenotyping is nec-
essary for whole- plant resistance breeding (Chuberre et al., 2018;
Tracy et al., 2020; Wiley et al., 2020). Rhizosphere phenotyping
opens new opportunities for experimental approaches, includ-
ing stress treatments, repeatability and combined use of imaging
techniques and machine learning to extract new traits from images,
within a systems approach (Tracy et al., 2020). The belowground net
primary production accounts for 40%– 70% of total terrestrial pro-
ductivity (Gherardi et al., 2020); therefore, more studies are needed
to explore response s of tree roots to climate and pol lution and quan-
tify root losses to belowground herbivores.
9.7 | Modeling and risk assessment
Model diversity constitutes a multi- purpose toolkit that can help
society to face the future challenges. Improving and enhancing
scientific communication in forest modeling is required as part of
this enterprise. The development of models integrating air pollution
and climate change data from long- term monitoring programs are
needed to improve forest research assessing interactions between
air pollution and climate change from the individual level to the stand
level. Future challenges include understanding of (i) the impacts of
air pollution on soil chemistry, (ii) the effects of climate change and
air pollution on plant phenol og y and reproductive fitness, (iii) the ca-
pacity of forests to sequester carbon under changing, and extremes,
climatic conditions and co- exposure to elevated levels of pollution,
and (iv) the effects of plant competitiveness (monocultures vs. mixed
cultures, single trees vs. community responses) on plant responses
to stressors.
ACKNOWLEDGMENTS
This work was outlined in the framework of the Research Group
8.04.00 “Air Pollution and Climate Change” under the International
Union of Forest Research Organizations (IUFRO). IUFRO is the
largest international network of forest scientists, promoting global
cooperation in forest- related research and enhancing the under-
standing of the ecological, economic, and social aspects of forests
and trees. M.V.K. was supported by the Academy of Finland (pro-
jects 276671, 311929, and 316182). M.W. was supported by JST
SICORP (JPMJSC16HB). A.D.M., P.S., Y.H., and E.P. were supported
by the LIFE projects MODERN (LIFE20 GIE/IT/000091), MOTTLES
(LIFE15 ENV/IT/000183), and AIRFRESH (LIFE19 ENV/FR/000086).
Contributions of JCB were partially supported through funding from
the US Department of Energy under award number DE- EM0005228
to the University of Georgia Research Foundation. Open Access
   
|
15
DE MARCO et al.
Funding provided by ENEA Agenzia Nazionale per Le Nuove
Tecnologie l'Energia e lo Sviluppo Economico Sostenibile within the
CRUI- CARE Agreement.
CONFLICT OF INTEREST
The authors declare no conflict of interest.
DATA AVA ILAB ILITY STATE MEN T
Dat a sharing is not applicable to this article as no new dat a were cre-
ated or analyzed in this study.
ORCID
Alessandra De Marco https://orcid.org/0000-0001-7200-2257
Alessio Collalti https://orcid.org/0000-0002-4980-8487
Enzai Du https://orcid.org/0000-0002-5519-0150
Mikhail V. Kozlov https://orcid.org/0000-0002-9500-4244
Elena Paoletti https://orcid.org/0000-0001-5324-7769
REFERENCES
Aber, J. D., Nadelhoffer, K. J., Steudler, P., & Melillo, J. M. (1989). Nitrogen
saturation in Northern forest ecosystems. BioScience, 39(6),
378– 386.
Agathokleous, E., Belz, R. G., Calatayud, V., De Marco, A., Hoshika, Y.,
Kitao, M., Saitanis, C. J., Sicard, P., Paoletti, E., & Calabrese, E. J.
(2019). Predicting the effect of ozone on vegetation via the linear
non- threshold (LNT), threshold and hormetic dose- response mod-
els. Science of the Total Environment, 6 49, 6174.
Agathokleous, E., Feng, Z., Oksanen, E., Sicard, P., Wang, Q., Saitanis,
C. J., Araminiene, V., Blande, J. D., Hayes, F., Calatayud, V.,
Domingos, M., Veresoglou, S. D., Peñuelas, J., Wardle, D. A ., De
Marco, A., Li, Z., Harmens, H., Yuan, X., Vitale, M., & Paoletti, E.
(2020). Ozone affects plant, insect, and soil microbial communi-
ties: A threat to terrestrial ecosystems and biodiversity. Scie nce
Advances, 6, eabc1176.
Agathokleous, E., Saitanis, C. J., Satoh, F., & Koike, T. (2015). Wild
plant species as subjects in O3 research. Eurasian Journal of Fores t
Research, 18, 1– 36.
Agathokleous, E., Saitanis, C. J., Wang, X., Watanabe, M., & Koike, T.
(2016). A review study on past 40 years of research on effects of
tropospheric O3 on belowground structure, functioning, and pro-
cesses of trees: A linkage with potential ecological implications.
Water, Air, & Soil Pollution, 227, 33.
Akselsson, C., Olsson, J., Belyazid, S., & Capell, R. (2016). Can increased
weathering rates due to future warming compensate for base
cation losses following whole- tree harvesting in spruce forests?
Biogeochemistry, 128, 89– 105.
Alonso, R., Elvira, S., González- Fernández, I., Calvete, H., García- Gómez,
H., & Bermejo, V. (2014). Drought stress does not protect Quercus
ilex L. from ozone effects: Results from a comparative study of two
subspecies differing in ozone sensitivit y. Plant Biology, 16, 375– 384.
Alonso, R., Elvira, S., Inclán, R., Bermejo, V., Castillo, F. J., & Gimeno, B.
S. (2003). Responses of Aleppo pine to ozone. In D. F. Karnosky,
K. E. Percy, A . H. Chappelka, & C. J. Simpson (Eds.), Air pollution,
global change and forests in the new millenium (pp. 211– 230). Els evier
Science Ltd.
Anav, A., De Marco, A., Collalti, A ., Emberson, L., Feng, Z., Lombardozzi,
D., Sicard, P., Verbeke, T., Viovy, N., Vitale, M., & Paoletti, E. (2022).
Legislative and functional aspects of different metrics used for
ozone risk assessment to forests. Environmental Pollution, 295,
118690.
Anav, A., De Marco, A., Proietti, C., Alessandri, A., Dell'Aquila, A., Cionni,
I., Friedlingstein, P., Khvorostyanov, D., Menut, L., Paoletti, E.,
Sicard, P., Sitch, S., & Vitale, M. (2016). Comparing concentration-
based (AOT40) and stomatal uptake (PODY) metrics for ozone
risk assessment to European forests. Global Change Biology, 22,
16081627.
Antoniadis, V., Levizou, E., Shaheen, S. M., Ok, Y. S., Sebastian, A., Baum,
C., Prasad, M. N. V., Wenzel, W. W., & Rinklebe, J. (2017). Trace
elements in the soil- plant interface: Phytoavailability, transloca-
tion, and phytoremediation– A review. Earth- Science Reviews, 171,
621– 645.
Aoyama, M., Hirose, K., & Igarashi, Y. (2006). Re- construction and up-
dating our understanding on the global weapons tests 137Cs fallout.
Journal of Environmental Monitoring, 8, 431.
Archer, C. R., Pirk, C. W. W., Carvalheiro, L. G., & Nicolson, S. W.
(2014). Economic and ecological implications of geographic bias
in pollinator ecology in the light of pollinator declines. Oikos, 123,
401– 407.
Augustaitis, A., Augustaitienė, I., Kliučius, A., Pivoras, G., Šopauskienė,
D., & Girgždienė, R. (2010). The seasonal variability of air pollu-
tion effects on pine conditions under changing climates. European
Journal of Forest Researc, 129, 431– 441.
Augustaitis, A., & Bytnerowicz, A. (2008). Contribution of ambient
ozone to Scots pine defoliation and reduced growth in the Central
European forests: A Lithuanian case stud. Environmental Pollution,
155, 436– 445.
Badea, O., Silaghi, D., Taut, I., Neagu, S., & Leca, S. (2013). Forest
monitoring- assessment, analysis and warning system for forest
ecosystem status. Notulae Botanicae Horti Agrobotanici Cluj- Napoca,
41, 613– 625.
Baig, S., Medlyn, B. E., Mercado, L., & Zaehle, S. (2015). Does the growth
response of woody plants to elevated CO2 increase with tempera-
ture? A model- oriented meta- analysis. Global Change Biology, 21,
4303– 4319.
Bebber, D. P. (2021). The gap between atmospheric nitrogen deposi-
tion experiments and reality. Science of the Total Environment, 801,
149774.
Belyazid, S., Phelan, J., Nihlgård, B., Sverdrup, H., Driscoll, C., Fernandez,
I., Aherne, J., Teeling- Adams, L. M., Bailey, S., Arsenault, M.,
Cleavitt, N., Engstrom, B., Dennis, R., Sperduto, D., Werier, D., &
Clark, C. (2019). Assessing the effects of climate change and air
pollution on soil properties and plant diversity in northeastern U.S.
hardwood forests: Model setup and evaluation. Water, Air, & Soil
Pollution, 230, 1– 33.
Belyazid, S., Sverdrup, H., Kurz, D., & Braun, S. (2010). Exploring ground
vegetation change for different deposition scenarios and methods
for estimating critical loads for biodiversity using the for SAFE- VEG
model in Switzerland and Sweden. Water, Air, & Soil Pollution, 216 ,
2 8 9 3 1 7 .
Benjamin, M. T., & Winer, A. M. (1998). Estimating the ozone e forming
potential of urban trees and shrubs. Atmospheric Environment, 32,
5 3 6 8 .
Beresford, N. A., Fesenko, S., Konoplev, A., Skuterud, L., Smith, J. T., &
Voigt, G. (2016). Thirty years after the Chernobyl accident: What
lessons have we learnt? Journal of Environmental Radioactivity, 157,
7 7 8 9 .
Beresford, N. A ., Scott, E. M., & Copplestone, D. (2020). Field effects
studies in the Chernobyl Exclusion Zone: Lessons to be learnt.
Journal of Environmental Radioactivity, 211, 105893.
Berger, T. W., Türtscher, S., Berger, P., & Lindebner, L. (2016). A slight
recovery of soils from acid rain over th e last thre e dec ade s is not re-
flected in the macro nutrition of beech (Fagus sylvatica) at 97 forest
stands of the Vienna woods. Environmental Pollution, 216, 6 2 4 6 3 5 .
Blanco, J. A. (2013). Modelos ecológicos: Descripción, explicación y pre-
dicción. Ecosistemas, 22(3), 1– 5.
16 
|
    DE MARCO et al.
Blan co, J. A., Ameztegui, A., & Rodríg uez, F. (2020) . Mod ellin g forest eco -
systems: A crossroad between scales, techniques and applications.
Ecological Modelling, 425, 109030.
Blande, J. D. (2021). Effects of air pollution on plant- insect interac-
tions mediated by olfactory and visual cues. Current Opinion in
Environmental Science & Health, 19, 100228.
Boccuzzi, G., Nakazato, R. K., Pereira, M. A. G., Rinaldi, M. C. S., Lopes,
M. I. M. S., & Domingos, M. (2021). Anthropogenic deposition in-
creases nitrogen- phosphorus imbalances in tree vegetation, litter
and soil of Atlantic Forest remnants. Plant and Soil, 461, 341– 354.
Bonan, G. B. (2008). Forests and climate change: Forcings, feedbacks,
and the climate benefits of forests. Science, 320, 1444– 1449.
Bruckman, V. J., Terada, T., Fukuda, K., Yamamoto, H., & Hochbichler, E.
(2016). Overmature periurban QuercusCarpinus coppice forests in
Austria and Japan: A comparison of carbon stocks, stand charac-
teristics and conversion to high forest. European Journal of Forest
Research, 135, 857– 869.
Butler, T., Vermeylen, F., Lehmann, C. M., Likens, G. E., & Puchalski, M.
(2016). Increasing ammonia concentration trends in large regions
of the USA derived from the NADP/AMoN network. Atmospheric
Environment, 146, 132– 140.
Bytnerowicz, A., Hsu, Y. M., Percy, K., Legge, A., Fenn, M. E., Schilling,
S., Frączek, W., & Alexander, D. (2016). Ground- level air pollution
changes during a boreal wildland mega- fire. Science of the Total
Environment, 572, 755– 769.
Calfapietra, C., Fares, S., Manes, F., Morani, A., Sgrigna, G., & Loreto,
F. (2013). Role of biogenic volatile organic compounds (BVOC)
emitted by urban trees on ozone concentration in cities: A review.
Environmental Pollution, 183, 71– 80.
Cassimiro, J. C., Moura, B. B., Alonso, R., Meirelles, S. T., & Moraes, R. M.
(2016). Ozone stomatal flux and O3 concentration- based metrics
for Astronium graveolens Jacq., a Bra zi li an native forest tr ee species.
Environmental Pollution, 213, 1007– 1015.
Cazzolla Gatti, R., Reich, P. B., Gamarra, J. G. P., Crowther, T., Hui, C.,
Morera, A., Bastin, F., de Miguel, S., Nabuurs, G.- J., Svenning, J.-
C., Serra- Diaz, J. M., Merow, C., Enquist, B., Kamenetsky, M., Lee,
J., Zhu, J., Fang, J., Jacobs, D. F., Pijanowski, B., Liang, J. (2022).
The number of tree species on Earth. Proceedings of the National
Academy of Sciences of the United States of America, e2115329119.
https://doi.or g/10.1073/pnas.21153 29119
Chino, M., Nakayama, H., Nagai, H., Terada, H., Katata, G., & Yamazawa,
H. (2011). Preliminary estimation of release amounts of 131I and
137Cs accidentally discharged from the Fukushima Daiichi Nuclear
Power Plant into the atmosphere. Journal of Nuclear Science and
Technology, 48(7), 1129– 1134.
Chiwa, M. (2020). Ten- year determination of atmospheric phospho-
rus deposition at three forested sites in Japan. Atmospheric
Environment, 223, 117247.
Chuberre, C., Plancot, B., Driouich, A., Moore, J. P., Bardor, M., Gügi, B.,
& Vicré, M. (2018). Plant immunity is compartmentalized and spe-
cialized in roots. Frontiers in Plant Science, 9, 1692.
Clark, C. M., Bai, Y., Bowman, W. D., Cowles, J. M., Fenn, M. E., Gilliam,
F. S., Phoenix, G. K., Siddique, I., Stevens, C. J., Sverdrup, H.
U., & Throop, H. L. (2013). Nitrogen Deposition and Terrestrial
Biodiversity. In S. A. Levin (Ed.), Encyclopedia of biodiversity (2nd ed.,
pp. 519– 536). Elsevier Inc.
Collalti, A., Ibrom, A., Stockmarr, A., Cescatti, A., Alkama, R., Fernández-
Martínez, M., Ciais, P., Sitch, S., Friedlingstein, P., Goll, D. S., Nabel,
J. E. M. S., Pongratz, J., Arneth, A., Haverd, V., & Prentice, I. C.
(2020). Forest production efficiency increases with growth tem-
perature. Nature Communications, 11, 5322.
Collalti, A ., Marconi, S., Ibrom, A., Trotta, C., Anav, A., D'Andrea,
E., Matteucci, G., Montagnani, L., Gielen, B., Mammarella, I.,
Grünwald, T., Knohl, A., Berninger, F., Zhao, Y., Valentini, R., &
Santini, M. (2016). Validation of 3D- CMCC Forest Ecosystem
Model (v.5.1) against eddy covariance data for 10 European forest
sites. Geoscientific Model Development, 9, 479– 504.
Collalti, A., Trotta, C., Keenan, T., Ibrom, A., Bond- Lamberty, B., Grote, R.,
Vicca, S., Reyer, C. P. O., Migliavacca, M., Veroustraete, F., Anav, A.,
Campioli, M., Scoccimarro, E., Grieco, E., Cescatti, A., & Matteucci,
G. (2018). Thinning can reduce losses in carbon use efficiency and
carbon stocks in managed forests under warmer climate. Journal of
Advances in Modelling Earth Systems, 10(10), 2427– 2452.
Committee on the Biological Effects of Ionizing Radiation. (1990). Health
effects of exposure to low levels of ionizing radiation: BEIR V. National
Research Council, National Academy Press.
Cooper, O. R., Parrish, D. D., Ziemke, J., Balashov, N. V., Cupeiro, M.,
Galbally, I. E., Gilge, S., Horowitz, L., Jensen, N. R., Lamarque, J.- F.,
Naik, V., Oltmans, S. J., Schwab, J., Shindell, D. T., Thompson, A. M.,
Thouret, V., Wang, Y., & Zbinden, R. M. (2014). Global distribution
and trends of tropospheric ozone: An observation- based review.
Elementa: Science of the Anthropocene, 2, 000029.
Coordination Centre for Effects. (2017). European critical loads:
Database , biodiversity and ecosystems at risk. CCE final report 2017
(J.- P. Hettelingh, M. Posch, & J. Slootweg, Eds.). Rijksinstituut voor
Volksgezondheid en Milieu.
Cunha- Zeri, G., & Ometto, J. (2021). Nitrogen emissions in Latin America:
A conceptual framework of drivers, impacts, and policy responses.
Environmental Development, 38, 100605.
Curran, P. J., Dungan, J. L., & Gholz, H. L. (1992). Seasonal LAI in slash
pine estimated with Landsat TM. Remote Sensing of Environment,
39(1), 3– 13.
Curtis, P. S., & Wang, X. Z. (1998). A meta- analysis of elevated CO2 ef-
fects on woody plant mass, form, and physiology. Oecologia, 113,
299– 313.
Dale, V. H., Joyce, L. A., McNulty, S., & Neilson, R. P. (2000). The interplay
between climate change, forests, and disturbances. Science of the
Total Environment, 262, 201204.
De Jager, N. R., Drohan, P. J., Miranda, B. M., Sturtevant, B. R., Stout, S. L.,
Royo, A. A., Gustafson, E. J., & Romanski, M. C. (2017). Simulating
ungulate herbivory across forest landscapes: A browsing extension
for LANDIS- II. Ecological Modelling, 350, 11– 2 9.
De Marco, A., Anav, A., Sicard, P., Feng, Z., & Paoletti, E. (2020). High
spatial resolution ozone risk assessment for Asian forests.
Environmental Research Letters, 15, 104095.
De Marco, A., Screpanti, A., Attorre, F., Proietti, C., & Vitale, M. (2013).
Assessing ozone and nitrogen impact on net primary productivity
with a generalised non- linear model. Environmental Pollution, 172,
250– 263.
De Marco, A., & Sicard, P. (2019). Why do we still need to derive ozone
critical levels for vegetation protection? Opinion paperIJESNR
21— October 2019.
De Marco, A., Sicard, P., Vitale, M., Carriero, G., Renou, C., & Paoletti, E.
(2015). Metrics of ozone risk assessment for Southern European
forests: Canopy moisture content as a potential plant response in-
dicator. Atmospheric Environment, 120, 182– 190.
de Vries, W. (2021). Impacts of nitrogen emissions on ecosystems and
health: A mini review. Current Opinion in Environmental Science &
Health, 19, 100249.
De Vries, W., Wamelink, G. W. W., van Dobben, H., Kros, J., Reinds, G. J.,
Mol- Dijkstra, J. P., Smart, S. M., Evans, C. D., Rowe, E. C., Belyazid,
S., Sverdrup, H. U., van Hinsberg, A., Posch, M., Hettelingh, J.- P.,
Spranger, T., & Bobbink, R. (2010). Use of dynamic soil– vegetation
models to assess impacts of nitrogen deposition on plant species
composition: An overview. Ecological Applications, 20, 60– 79.
Deepak, M., Lihavainen, J., Keski- Saari, S., Kontunen- Soppela, S.,
Tenkanen, A., Oksanen, E., & Kei nen, M. (2018). Genotypic varia-
tion and provenance- related clinal patterns in leaf surface second-
ary compounds of silver birch. Canadian Journal of Forest Research,
48, 494– 505.
   
|
17
DE MARCO et al.
Deryabina, T. G., Kuchmel, S. V., Nagorskaya, L. L., Hinton, T. G., Beasley,
J. C., Lerebours, A., & Smith, J. T. (2015). Long term census data
reveal abundant wildlife populations at chernobyl. Current Biology,
25, R824– R826.
Dirnböck, T., Foldal, C., Djukic, I., Kobler, J., Haas, E., Kiese, R., & Kitzler,
B. (2017). Historic nitrogen deposition determines future climate
change effects on nitrogen retention in temperate forests. Climatic
Change, 144, 221– 235.
Dirnböck, T., Pröll, G., Austnes, K., Beloica, J., Beudert, B., Canullo, R.,
De Marco, A., Fornasier, M. F., Futter, M., Goergen, K., Grandin,
U., Holmberg, M., Lindroos, A.- J., Mirtl, M., Neir ynck, J., Pecka, T.,
Nieminen, T. M., Nordbakken, J.- F., Posch, M., … Forsius, M. (2018).
Currently legislated decreases in nitrogen deposition will yield only
limited plant species recovery in European forests. Environmental
Research Letters, 13, 125010.
Dobson, M. C., Ulaby, T., Le Toan, T., Beaudoin, A., & Kasischke, E. S.
(1992). Dependence of radar backscatter on coniferous forest
biomass. IEEE Transactions on Geoscience and Remote Sensing, 30,
412– 41 5.
Domingos, M., Klumpp, A., & Klumpp, G. (2003). Disturbances to the
Atlantic rain forest in southeast Brazil. In E. Emberson, M. Ashmore,
& F. Murray (Eds.), Air pollution impacts on vegetation in developing
countries (Vol. 1a, pp. 287– 308). Imperial College Press.
Du, E. (2016). Rise and fall of nitrogen deposition in the United States.
Proceedings of the National Academy of Sciences of the United States
of America, 113, E3594– E3595.
Du, E., & de Vries, W. (2018). Nitrogen- induced new net primary pro-
duction and carbon se- questration in global forests. Environmental
Pollution, 242, 1476– 1487.
Du, E., De Vries, W., Han, W., Liu, X., Yan, Z., & Jiang, Y. (2016). Imbalanced
phosphorus and nitrogen deposition in China's forests. Atmospheric
Chemistry and Physics, 16 , 8571– 8579.
Du, E., De Vries, W., McNulty, S., & Fenn, M. E. (2018). Bulk deposition of
base cationic nutrients in China's forests: Annual rates and spatial
characteristics. Atmospheric Environment, 184, 121– 128.
Du, E., Dong, D., Zeng, X., Sun, Z., Jiang, X., & de Vries, W. (2017). Direct
effect of acid rain on leaf chlorophyll content of terrestrial plants in
China. Science of the Total Environment, 605, 76 4– 769.
Du, E., Terrer, C., Pellegrini, A . F. A., Ahlström, A ., van Lissa, C. J., Zhao,
X., Xia, N., Wu, X., & Jackson, R. B. (2020). Global patterns of ter-
restrial nitrogen and phosphorus limitation. Nature Geoscience, 13,
221– 226.
Duan, L., Ma, X., Larssen, T., Mulder, J., & Hao, J. (2011). Response of
surface water acidification in Upper Yangtze River to SO2 emis-
sions abatement in China. Environmental Science & Technology, 45,
3275– 3281.
Duan, L., Yu, Q., Zhang, Q., Wang, Z., Pan, Y., Larssen, T., Tang, J., &
Mulder, J. (2016). Acid deposition in Asia: Emissions, deposition,
and ecosystem effects. Atmospheric Environment, 146, 55– 69.
Dubey, A., Malla, M. A., Khan, F., Chowdhary, K., Yadav, S., Kumar, A.,
Sharma, S., Khare, P. K., & Khan, M. L. (2019). Soil microbiome: A
key player for conservation of soil health under changing climate.
Biodiversity and Conservation, 28(8– 9), 2405– 2429.
Dungey, H. S., Dash, J. P., Pont, D., Clinton, P. W., Watt, M. S., & Telfer,
E. J. (2018). Phenotyping whole forests will help to track genetic
performance. Trends in Plant Science, 23(10), 854– 864.
Emberson, L., Ashmore, M. R., Cambridge, H. M., Simpson, D., &
Tuovinen, J. P. (2000). Modelling stomatal ozone flux across Europe.
Environmental Pollution, 109, 403– 413.
Engardt, M., Simpson, D., Schwikowski, M., & Granat, L. (2017).
Deposition of sulphur and nitrogen in Europe 1900– 2050. Model
calculations and comparison to historical observations. Tellus B:
Chemical and Physical Meteorology, 69(1), S. 1328945.
Engela, M. R. G. S., Furlan, C. M., Esposito, M. P., Fernandes, F. F., Carrari,
E., Domingos, M., Paoletti, E., & Hoshika, Y. (2021). Metabolic
and physiological alterations indicate that the tropical broadleaf
tree Eugenia uniflora L. is sensitive to ozone. Science of the Total
Environment, 769, 145080.
Eränen, J. K., Nielsen, J., Zverev, V. E., & Kozlov, M. V. (2009). Mountain
birch under multiple stressors— Heavy metal resistant popula-
tions co- resistant to biotic stress but maladapted to abiotic stress.
Journal of Evolutionary Biology, 22, 840851.
Etzold, S., Ferretti, M., Reinds, G. J., Solberg, S., Gessler, A., Waldner, P.,
Schaub, M., Simpson, D., Benham, S., Hansen, K., & Ingerslev, M.
(2020). Nitrogen deposition is the most important environmental
driver of growth of pure, even- aged and managed European for-
ests. Forest Ecology and Management, 458, 117762.
European Commission. (2020). Communication from the Commission to
the European parliament, the council, the European economic and so-
cial committee and the committee of the regions. EU biodiversity strat-
egy for 2030— Bringing nature back into our lives. COM (2020) 380
final. Brussels, Belgium.
Evans, M. R. (2012). Modelling ecological systems in a changing world.
Philosophical Transac tions of the Royal Societ y B: Biological Sciences,
367, 181190.
Fabrika, M., Valent, P., & Merganičová, K. (2019). Forest modelling and
visualisation— State of the art and perspectives. Central European
Forestry Journal, 65, 147 165.
Falk, T., Herndon, N., Grau, E., Buehler, S., Richter, P., Zaman, S., Baker,
E. M., Ramnath, R., Ficklin, S., Staton, M., & Feltus, F. A. (2018).
Growing and cultivating the forest genomics database, TreeGenes.
Database, 2018, bay084.
Fang, Y., Koba, K., Makabe, A., Takahashi, C., Zhu, W., Hayashi, T., Hokari,
A., Urakawa, R., Bai, E., Houlton, B. Z., & Xi, D. (2015). Microbial
denitrification dominates nitrate losses from forest ecosystems.
Proceedings of the National Academy of Sciences of the United States
of America, 112, 1470– 1474.
Feng, Z., Shang, B., Gao, F., & Calatayud, V. (2019). Current ambient
and elevated ozone effects on poplar: A global meta- analysis
and response relationships. Science of the Total Environment, 654,
832840.
Feng, Z., Shang, B., Li, Z., Calatayud, V., & Agathokleous, E. (2019). Ozone
will remain a threat for plants independently of nitrogen load.
Functional Ecology, 33, 1854– 1870.
Feng, Z., Yuan, X., Fares, S., Loreto, F., Li, P., Hoshika, Y., & Paoletti, E.
(2019). Isoprene is more affected by climate drivers than monoter-
penes: A meta- analytic review on plant isoprenoid emissions. Plant,
Cell & Environment, 42, 1939– 1949.
Fernandes, F. F., Esposito, M. P., Engela, M. R. G. S., Cardoso- Gustavson,
P., Furlan, C. M., Hoshika, Y., Carrari, E., Magni, G., Domingos, M.,
& Paoletti, E. (2019). The passion fruit liana (Passiflora edulis Sims,
Passifloraceae) is tolerant to ozone. Science of the Total Envi ronment,
656, 1091– 1101.
Fischer, R., Aas, W., de Vries, W., Clarke, N., Cudlin, P., Leaver, D., Lundin,
L., Matteucci, G., Matyssek, R., Mikkelsen, T. N., Mirtl, M., Öztürk,
Y., Papale, D., Potocic, N., Simpson, D., Tuovinen, J. P., Vesala, T.,
Wieser, G., & Paoletti, E. (2011). Towards a transnational system
of supersites for forest monitoring and research in Europe— An
overview on present state and future recommendations. iFore st-
Biogeosciences and Forestr y, 4(4), 167.
Flannigan, M., Stocks, B., Turetsky, M., & Wotton, M. (2009). Impacts of
climate change on fire activity and fire management in the circum-
boreal forest. Global Change Biology, 15(3), 49– 560.
Fleck, S., Ahrends, B., Sutmöller, J., Albert, M., Evers, J., & Meesenburg,
H. (2017). Is biomass accumulation in forests an option to prevent
climate change induced increases in nitrate concentrations in the
north German lowland? Forests, 8, 219.
Forsius, M., Kujala, H., Minunno, F., Holmberg, M., Leikola, N., Mikkonen,
N., Autio, I., Paunu, V.- V., Tanhuanpää, T., Hurskainen, P., Mäyrä, J.,
Kivinen, S., Keski- Saari, S., Kosenius, A.- K., Kuusela, S., Virkkala, R.,
Viinikka, A ., Vihervaara, P., Akujärvi, A., … Heikkinen, R. K . (2021).
Developing a spatiallyexplicit modelling and evaluation framework
18 
|
    DE MARCO et al.
for integrated carbon sequestrationand biodiversity conservation:
Application in southern Finland. Science of the Total Environment,
775, 145847.
Frank, D., Reichstein, M., Bahn, M., Thonicke, K., Frank, D., Mahecha,
M. D., Smith, P., Van der Velde, M., Vicca, S., Babst, F., & Beer, C.
(2015). Effects of climate extremes on the terrestrial carbon cycle:
Concepts, processes and potential future impacts. Global Change
Biology, 21(8), 2861– 2880.
Fukuyama, T., Onda, Y., Takenaka, C., & Walling, D. E. (2008). Investigating
erosion rates within a Japanese cypress plantation using Cs- 137
and Pb- 210 ex measurements. Journal of Geophysical Research, 113,
F02007.
Gao, F., Catalayud, V., Paoletti, E., Hoshika, Y., & Feng, Z. (2017). Water
stress mitigates the negative effects of ozone on photosynthesis
and biomass in poplar plants. Environmental Pollution, 230, 268– 279.
Gao, Y., Ma, M., Yang, T., Chen, W., & Yang, T. (2018). Global atmospheric
sulfur deposition and associated impaction on nitrogen cycling in
ecosystems. Journal of Cleaner Production, 195, 1– 9.
Garmo, O. A., Skjelkvale, B. L ., De Wit, H. A., Colombo, L., Curtis, C.,
Fölster, J., Hoffmann, A., Hruška, J., Jef fries DS, H. T., Keller, W.
B., Krám, P., Majer, V., Monteith, D. T., Paterson, A. M., Rogora,
M., Rzychon, D., Steingruber, S., Stoddard, J. L., Vuorenmaa, J., &
Worsztynowicz, A. (2014). Trends in surface water chemistr y in
acidified areas in Europe and North America from 1990 to 2008.
Water Air and Soil Pollution, 225, 1880.
Gaudio, N., Belyazid, S., Gendre, X., Mansat, A., Nicolas, M., Rizzetto, S.,
Sverdrup, H., & Probst, A. (2015). Combined effect of atmospheric
nitrogen deposition and climate change on temperate forest soil
biogeochemistry: A modeling approach. Ecological Modelling, 306,
24– 34 .
Gauthier, S., Bernier, P., Kuuluvainen, T., Shvidenko, A. Z., &
Schepaschenko, D. G. (2015). Boreal forest health and global
change. Science, 349(6250), 819– 822.
Gawel, J. E., Ahner, B. A., Friedland, A. J., & Morel, F. M. M. (1996). Role
for heav y metals in forest decline indicated by phytochelatin mea-
surements. Nature, 381(6577), 64– 65.
Geng, F., Tie, X., Guenther, A ., Li, G., Cao, J., & Harley, P. (2011). Effect
of isoprene emissions from major forests on ozone formation in
the city of Shanghai, China. Atmospheric Chemistry and Physics, 11,
104 49– 10459.
Gerken, T., Wei, D., Chase, R. J., Fuentes, J. D., Schumacher, C., Machado,
L. A. T., Andreoli, R. V., Chamecki, M., Souza, R. A. F., Freire, L. S.,
Jardine, A. B., Manzi, A. O., Santos, R. M. N., Randow, C., Costa, P.
S., Stoy, P. C., & Tóta, J. (2016). Trowbridge AM downward transport
of ozone rich air and implications for atmospheric chemistry in the
Amazon rainforest. Atmospheric Environment, 124, 64– 76.
Gherardi, L. A., & Sala, O. E. (2020). Global patterns and climatic controls
of belowground net carbon fixation. Proceedings of the National
Academy of S ciences of the Unite d States of America, 117, 202006715.
Gilliam, F. S., Burns, D. A ., Driscoll, C. T., Frey, S. D., Lovett, G. M., &
Watmough, S. A. (2019). Decreased atmospheric nitrogen deposi-
tion in eastern North America: Predicted responses of forest eco-
systems. Environmental Pollution, 244, 560– 574.
Giordani, P., Calatayud, V., Stofer, S., Seidling, W., Granke, O., & Fischer,
R. (2014). Detecting the nitrogen critical loads on European forests
by means of epiphytic lichens. A signal- to- noise evaluation. Fores t
Ecology and Management, 311, 29– 40.
Godzik, B. (2020). Use of bioindication methods in national, regional and
local monitoring in Poland- changes in the air pollution level over
several decades. Atmosphere, 11, 143.
Grams, T. E. E., Anegg, S., Häberle, K.- H., Langebartels, C., & Matyssek,
R. (1999). Interactions of chronic exposure to elevated CO2 and O3
levels in the photosynthetic light and dark reactions of European
beech (Fagus sylvatica). New Phytologist, 14 4, 95– 107.
Grantz, D. A., Garner, J. H. B., & Johnson, D. W. (2003). Ecological effects
of particulate matter. Environment International, 29, 21 3239.
Grantz, D. A., Gunn, S., & Vu, H.- B. (2006). O3 impact s on plant develop-
ment: A meta- analysis of root/shoot allocation and growth. Plant,
Cell Environ., 29, 1193– 1209.
Gray, C. M., Monson, R. K., & Fierer, N. (2010). Emissions of volatile or-
ganic compounds during the decomposition of plant litter. Journal
of Geophysical Research, 115, G03015.
Gustafson, E. J. (2013). When relationships estimated in the past cannot
be used to predict the future: Using mechanistic models to pre-
dict landscape ecological dynamics in a changing world. Landscape
Ecology, 28, 1429– 1437.
Hansen, K., Personne, E., Skjøth, C. A ., Loubet, B., Ibrom, A., Jensen, R.,
Sørensen, L. L., & Bøgh, E. (2017). Investigating sources of measured
forest- atmosphere ammonia fluxes using two- layer bi- directional
modelling. Agricultural and Forest Meteorology, 237, 80– 94.
Hansen, W. D., & Turner, M. G. (2019). Origins of abrupt change? Postfire
subalpine conifer regeneration declines nonlinearly with warming
and drying. Ecological Monographs, 89, e01340.
Hao, Z., Hao, F., Singh, V. P., & Zhang, X. (2018). Changes in the sever-
ity of compound drought and hot extremes over global land areas.
Environmental Research Letters, 13, 124022.
Harrell, P. A ., Bourgeau- Chavez, L. L., Kasischke, E. S., French, N. H. F.,
& Christensen, N. L. (1995). Sensitivity of ERS- 1 and JERS- 1 radar
data to biomass and stand structure in Alaskan boreal forest.
Remote Sensing of Environment, 54, 247260.
Hartmann, H., Schuldt, B., Sanders, T. G., Macinnis- Ng, C., Boehmer, H.
J., Allen, C. D., Bolte, A., Crowther, T. W., Hansen, M. C., Medlyn, B.
E., & Ruehr, N. K. (2018). Monitoring global tree mortality patterns
and trends. Report from the V W symposium Crossing scales and
disciplines to identify global trends of tree mortality as indicators
of forest health. New Phytologist, 217, 984– 987.
Hashimoto, S., Imamura, N., Kaneko, S., Komatsu, M., Matsuura, T.,
Nishina, K., & Ohashi, S. (2020). New predictions of 137Cs dynamics
in forests after the Fukushima nuclear accident. Scientific Reports,
10, 29.
Haukioja, E., Hanhimäki, S., & Walter, G. H. (1994). Can we learn about
herbivory on eucalypts from research on birches, or how general
are general plant- herbivore theories. Australian Journal of Ecology,
19, 1– 9.
He, M., He, C.- Q., & Ding, N.- Z. (2018). Abiotic stresses: General de-
fenses of land plants and chances for engineering multistress toler-
ance. Frontiers in Plant Science, 9, 1771.
Hein, L., White, L., Miles, A., & Roberts, P. (2018). Analysing the im-
pacts of air quality policies on ecosystem services; a case study
for Telemark, Norway. Journal of Environmental Management, 206,
650– 663.
Hellegers, M., Ozinga, W. A., Hinsberg, A., van Huijbregts, M. A. J.,
Hennekens, S. M., Schaminée, J. H. J., Dengler, J., & Schipper, A.
M. (2020). Evaluating the ecological realism of plant species dis-
tribution models with ecological indicator values. Ecography, 43,
16 1– 170 .
Hong, G. H., Hamilton, T. F., Baskaran, M., & Kenna, T. C. (2012).
Applications of anthropogenic radionuclides as tracers to in-
vestigate marine environmental processes. In M. Baskaran (Ed.),
Handbook of environmental isotope geochemistr y. advances in isotope
geochemistry (pp. 367– 394). Springer.
Hoshika, Y., Katata, G., Deushi, M., Watanabe, M., Koike, T., & Paoletti, E.
(2015). Ozone- induced stomatal sluggishness changes carbon and
water balance of temperate deciduous forests. Scientific Reports,
5, 1– 8.
Huang, C. Y., Anderegg, W. R., & Asner, G. P. (2019). Remote sensing of
forest die- off in the Anthropocene: From plant ecophysiology to
canopy structure. Remote Sensing of Environment, 231, 111233.
Huang, T., Zhu, X., Zhong, Q., Yun, X., Meng, W., Li, B., Ma, J., Zeng, E.
Y., & Tao, S. (2017). Spatial and temporal trends in global emissions
of nitrogen oxides from 1960 to 2014. Environmental Science &
Technology, 51, 7992– 8000.
   
|
19
DE MARCO et al.
Imamura, N., Komatsu, M., Ohashi, S., Hashimoto, S., Kajimoto, T.,
Kaneko, S., & Takano, T. (2017). Temporal changes in the ra-
diocesium distribution in forests over the five years after the
Fukushima Daiichi Nuclear Power Plant accident. Scientific
Reports, 7, 8179.
Imperato, V., Kowalkowski, L., Portillo- Estrada, M., Gawronski, S. W.,
Vangronsveld, J., & Thijs, S. (2019). Characterisation of the Carpinus
betulus L. phyllomicrobiome in urban and forest areas. Frontiers in
Microbiology, 10(1110). https://doi.org/10.3389/fmicb.2019.01110
Inness, A., Ribas, R., & Engelen, R. (2019). The use of Sentine l- 5P air qualit y
data by CAMS. ECMWF Newsletter No. 159, 25– 30.
Inter na ti on al Atomic Energy Agency. (20 06). Environmental consequences
of the chernobyl accident and their remediation: Twenty years of expe-
rience, radiological assessment reports series no. 8. IAEA .
Itoh, Y., Imaya, A., & Kobayashi, M. (2015). Initial radiocesium deposi-
tion on forest ecosystems surrounding the Tokyo metropolitan
area due to the Fukushima Daiichi Nuclear Power Plant accident.
Hydrological Research Letters, 9, 17.
IUCN. (2021). Forests and climate change. IUCN issues brief. h t tps://
www.iucn.org/sites/ dev/files/ fores ts_and_clima te_change_
issues_brief_2021.pdf
Izuta, T. (2017). Air pollution impacts on plants in East Asia. Springer.
Jakovljević, T., Lovreškov, L., Jelić, G., Anav, A., Popa, I., Fornasiere, M. F.,
Proietti, C., Limić, I., Butorac, L., Vitale, M., & De Marco, A. (2021).
Impact of ground- level ozone on Mediterranean forest ecosystems
health. Science of the Total Environment, 783, 147063.
Jansson, J. K., & Hofmockel, K. S. (2020). Soil microbiomes and climate
change. Nature Reviews Microbiology, 18(1), 35– 46.
Ji, K., Wang, Y., Du, L., Xu, C., Liu, Y., He, N., Wang, J., & Liu, Q. (2019).
Research progress on the biological effects of low- dose radiation in
China. Dose- Response, 17(1), 1559325819833488.
Johnson, J., Pannatier, E. G., Carnicelli, S., Cecchini, G., Clarke, N., Cools,
N., Hansen, K., Meesenburg, H., Nieminen, T. M., Pihl- Karlsson, G.,
Titeux, H., Vanguelova, E., Verstraeten, A., Vesterdal, L., Waldner,
P., & Jonard, M. (2018). The response of soil solution chemistr y
in European forests to decreasing acid deposition. Global Change
Biology, 24, 3603– 3619.
Kang, Y., Liu, M., Song, Y., Huang, X ., Yao, H., Cai, X., Zhang, H., Kang, L.,
Liu, X., Yan, X., He, H., Zhang, Q., Shao, M., & Zhu, T. (2016). High-
resolution ammonia emissions inventories in China from 1980 to
2012. Atmospheric Chemistr y and Physics, 16, 2043– 2058.
Karan, M., Liddell, M., Prober, S. M., Arndt, S., Beringer, J., Boer, M.,
Cleverly, J., Eamus, D., Grace, P., Van Gorsel, E., Hero, J.- M., Hutley,
L., Macfarlane, C., Metcalfe, D., Meyer, W., Pendall, E., Sebastian,
A., & Wardlaw, T. (2016). The Australian SuperSite network: A con-
tinental, long- term terrestrial ecosystem observatory. Science of the
Total Environment, 568, 1263– 1274.
Katanić, M., Paoletti, E., Orlović, S., Grebenc, T., & Kraigher, H. (2014).
Mycorrhizal status of an ozone- sensitive poplar clone treated
with the antiozonant ethylene diurea. European Journal of Forest
Research, 133(4), 735– 743.
Kato, H., Onda, Y., Saidin, Z. H., Sakashita, W., Hisadome, K., & Loffredo,
N. (2019). Six- year monitoring study of radiocesium transfer in for-
est environments following the Fukushima nuclear power plant ac-
cident. Journal of Environmental Radioactivity, 210, 105817.
Kefauver, S., Penuelas, J., & Ustin, S. (2012). Applications of hyperspec-
tral remote sensing and GIS for assessing forest health and air pollu-
tion. International Geoscience and Remote Sensing Symposium
(IGARSS), 3379– 3382.
Kimmins, J. P., Blanco, J. A., Seely, B., Welham, C., & Scoullar, K. (2008).
Complexity in modeling forest ecosystems; how much is enough?
Forest Ecology and Management, 256, 1646– 1658.
Kobayashi, R., Kobayashi, N. I., Tanoi, K., Masumori, M., & Tange, T. (2019).
Potassium supply reduces cesium uptake in Konara oak not by an
alteration of uptake mechanism, but by the uptake competition
between the ions. Journal of Environmental Radioactivity, 208– 209,
1060 32.
Koike, T., Kitao, M., Hikosaka, K., Agathokleous, E., Watanabe, Y.,
Watanabe, M., Eguchi, N., & Funada, R. (2018). Photosynthetic and
photosynthesis- related responses of Japanese native trees to CO2:
Results from phytotrons, open- top chambers, natural CO2 springs, and
free- air CO2 enrichment. Springer.
Kozlov, M. V. (2005). Pollution resistance of mountain birch, Betula pu-
bescens subsp. czerepanovii, near the copper- nickel smelter: Natural
selection or phenotypic acclimation? Chemosphere, 59, 189– 197.
Kozlov, M. V., Lanta, V., Zverev, V., & Zvereva, E. L. (2015). Background
losses of woody plant foliage to insects show variable relationships
with plant functional traits across the globe. Journal of Ecolog y, 103,
15191528.
Kozlov, M. V., Zverev, V., & Zvereva, E. L. (2017). Combined effects of en-
vironmental disturbance and climate warming on insect herbivory
in mountain birch in subarctic forests: Results of 26- year monitor-
ing. The Science of the Total Environment, 601- 602, 802– 811.
Kozlov, M. V., & Zvereva, E. L. (2015). Changes in the background losses
of woody plant foliage to insects during the past 60 years: Are the
predictions fulfilled? Biology Letters, 11, 20150480.
Kozlov, M. V., Zvereva, E. L., & Zverev, V. E. (2009). Impacts of point pollut-
ers on terre strial biota: Comp arative analysis of 18 conta minated areas.
Springer.
Krause, A., Pugh, T. A. M., Bayer, A. D., Doelman, J. C., Humpenöder, F.,
Anthoni, P., Olin, S., Bodirsky, B. L., Popp, A., Stehfest, E., & Arneth,
A. (2017). Global consequences of afforestation and bioenergy
cultivation on ecosystem service indicators. Biogeosciences, 14,
48294850.
Krüger, I., Sanders, T. G. M., Potočić, N., Ukonmaanaho, L., & Rautio, P.
(2020). Increased evid ence of nutrient imb alances in fores t trees across
Europe (ICP forests brief no. 4). Programme co- ordinating Centre of
ICP Forests, Thünen Institute of Forest Ecosystems.
Kumar, A., Bali, K., Singh, S., Naja, M., & Mishra, A. K. (2019). Estimates of
reactive trace gases (NMVOCs, CO and NOx) and their ozone form-
ing potentials during forest fire over Southern Himalayan region.
Atmospheric Research, 227, 41– 51.
Kurokawa, J., & Ohara, T. (2020). Long- term historical trends in air pol-
lutant emissions in Asia: Regional Emission inventory in ASia (REAS)
version 3. Atmospheric Chemistry and Physics, 20, 1276112793.
Lausch, A., Borg, E., Bumberger, J., Dietrich, P., Heurich, M., Huth, A.,
Jung, A., Klenke, R., Knapp, S., Mollenhauer, H., & Paasche, H.
(2018). Understanding forest health with remote sensing, part III:
Requirements for a scalable multi- source forest health monitoring
network based on data science approaches. Remote Sensing, 10(7),
1120 .
Le Toan, T., Beaudoin, A., & Guyon, D. (1992). Relating forest biomass to
SAR data. IEEE Transactions on Geoscience and Remote Sensing, 30,
40 3– 411.
Lechner, A. M., Foody, G. M., & Boyd, D. S. (2020). Applications in remote
sensing to forest ecology and management. One Earth, 2, 405– 412.
Lefohn, A. S., Malley, C. S., Smith, L., Wells, B., Hazucha, M., Simon, H.,
Naik, V., Mills, G., Schultz, M. G., Paoletti, E., & De Marco, A. (2018).
Tropospheric ozone assessment report: Global ozone metrics
for climate change, human health, and crop/ecosystem research.
Elementa: Science of the Anthropocene, 6, 28.
Li, C., McLinden, C., Fioletov, V., Krotkov, N., Carn, S., Joiner, J., Streets,
D., He, H., Ren, X., Li, Z., & Dickerson, R. R. (2017). India is over-
taking china as the world’s largest emitter of anthropogenic sulfur
dioxide. Scientific Reports, 7, 14304.
Li, P., De Marco, A., Feng, Z., Anav, A., Zhou, D., & Paoletti, E. (2018).
Nationwide ground- level ozone measurements in China suggest se-
rious risks to forests. Environmental Pollution, 237, 803– 813.
Li, P., Lin, C., Cheng, H., Duan, X., & Lei, K. (2015). Contamination and
health risks of soil heavy metals around a lead/zinc smelter in
20 
|
    DE MARCO et al.
southwestern China. Ecotoxicology and Environmental Safety, 113 ,
391– 399.
Li, P., Yin, R., Zhou, H., Yuan, X., & Feng, Z. (2021). Soil pH drives poplar
rhizosphere soil microbial community responses to ozone pollution
and nitrogen addition. European Journal of Soil Science, 73, 1– 14.
Lilleskov, E. A., Kuyper, T. W., Bidartondo, M. I., & Hobbie, E. A. (2019).
Atmospheric nitrogen deposition impacts on the structure and
function of forest mycorrhizal communities: A review. Environmental
Pollution, 246, 148– 162.
Lincoln, D. E., Fajer, E. D., & Johnson, R. H. (1993). Plant- insect herbi-
vore interactions in elevated CO2 environments. Trends Ecol Evol,
8, 64– 68.
Liu, L. L., & Greaver, T. L. (2009). A review of nitrogen enrichment effects
on three biogenic GHGs: The CO2 sink may be largely offset by
stimulated N2O and CH4 emission. Ecology Letters, 12, 1103– 1117.
Liu, W., Sun, F., Lim, W. H., Zhang, J., Wang, H., Shiogama, H., & Zhang,
Y. (2018). Global drought and severe drought- affected populations
in 1.5 and 2°C warmer worlds. Earth System Dynamics, 9, 267– 283.
Logan, J. A ., Regniere, J., & Powell, J. A. (2003). Assessing the impacts of
global warming on forest pest dynamics. Frontiers in Ecology and the
Environment, 1, 130– 137.
Lu, Z., Zhang, Q., & Streets, D. G. (2011). Sulfur dioxide and primary
carbonaceous aerosol emission in China and India, 1996- 2010.
Atmospheric Chemistry and Physics, 11 , 9839– 9864.
Ludovisi, R., Tauro, F., Salvati, R., Khoury, S., Mugnozza, G. S., &
Harfouche, A. (2017). UAV- based thermal imaging for high-
throughput field phenotyping of black poplar response to drought
front. Plant Science, 8, 1681.
Lundin, L., & Forsius, M. (2004). International cooperative programme
on integrated monitoring of air pollution effects on ecosystems
(ICP integrated monitoring). In J. Sliggers & W. Kakebeeke (Eds.),
Clearing the air: 25 years of the convention on long- range transbound-
ary air pollution (p. 70). United Nations Economic Commission for
Europe.
Lyons, P. C., Okuda, K., Hamilton, M., Hinton, T. G., & Beasley, J. C.
(2020). Rewilding of Fukushima's human evacuation zone. Frontiers
in Ecology and the Environment, 18, 127– 134.
Maas, R., & P. Grennfelt (Eds.). (2016). Towards cleaner air. Scientific as-
sessment report 2016. EMEP Steering Body and Working Group on
effects of the convention on long- range transboundary air pollu-
tion, Oslo, xx+50 pp.
Maass, M., Balvanera, P., Bourgeron, P., Equihua, M., Baudry, J., Dick, J.,
Forsius, M., Halada, L., Krauze, K., Nakaoka, M., Orenstein, D. E.,
Parr, T. W., Redman, C. L., Rozzi, R., Santos- Reis, M., Swemmer, A.
M., & Vădineanu, A. (2016). Changes in biodiversity and trade- offs
among ecosystem services, stakeholders, and components of well-
being: the contribution of the International LongTerm Ecological
Research network (ILTER) to Programme on Ecosystem Change and
Society (PECS). Ecology and Society, 21(3), 31.
Mac Nally, R., Cunningham, S. C., Baker, P. J., Horner, G. J., & Thomson,
J. R. (2011). Dynamics of Murray- Darling floodplain forests under
multiple stressors: The past, present, and future of an Australian
icon. Water Resources Research, 47, W00G05.
Maja, M., Kasurinen, A., Holopainen, T., Kontunen- Soppela, S., Oksanen,
E., & Holopainen, J. K. (2015). Volatile organic compounds emitted
from silver birch of different provenances across a latitudinal gradi-
ent in Finland. Tree Physiology, 35, 975– 986.
Mäki, M., Aaltonen, H., Heinonsalo, J., Hellén, H., Pumpanen, J., & Bäck,
J. (2019). Boreal forest soil is a significant and diverse source of
volatile organic compounds. Plant Soil, 441, 89– 110.
Manaka, T., Imamura, N., Kaneko, S., Miura, S., Furusawa, H., & Kanasashi,
T. (2019). Six- year trends in exchangeable radiocesium in Fukushima
forest soils. Journal of Environmental Radioactivity, 203, 84– 92.
Manninen, T., Stenberg, P., Rautiainen, M., Smolander, H., & Voipio,
P. (2003). Estimation of boreal forest LAI using C- band SAR.
In Environmental Science, Mathematics IGARSS 2003, 2003 IEEE
International Geoscience and Remote Sensing Symposium, Proceedings
(IEEE Cat. No. 03CH37477), pp. 1631– 1633.
Marano, G., Langella, G., Basile, A., Cona, F., De Michele, C., Manna, P.,
Teobaldelli, M., Saracino, A., & Terribile, F. A. (2019). Geospatial de-
cision support system tool for supporting integrated forest knowl-
edge at the landscape scale. Fores ts, 10 (8), 690.
Maréchaux, I., Langerwisch, F., Huth, A., Bugmann, H., Morin, X., Reyer,
C. P., Seidl, R., Collalti, A., Dantas de Paula, M., Fischer, R., & Gutsch,
M. (2020). Tackling unresolved questions in forest ecology: The
past and future role of simulation models. Ecology and Evolution,
2021(11), 3746– 3770.
Matyssek, R., Kozovits, A. R., Wieser, G., King, J., & Rennenberg, H.
(2017). Woody- plant ecosystems under climate change and air
pollution— Response consistencies across zonobiomes? Tree
Physiolog y, 37, 706– 732.
Matyssek, R., Wieser, G., Calfapietra, C ., De Vries, W., Dizengremel, P.,
Ernst, D., Jolivet, Y., Mikkelsen, T. N., Mohren, G. M. J., Le Thiec, D.,
Tuovinen, J. P., Weatherall, A., & Paoletti, E. (2012). Forests under
climate change and air pollution: Gaps in understanding and future
directions for research. Environmental Pollution, 160(1), 57– 65.
McDermott, A. (2020). News feature: Foreseeing fires. Proceeding s of the
National Academy of Sciences of the United States of America, 117,
21834– 21838.
Merganičová, K., Merganič, J., Lehtonen, A., Vacchiano, G., Zorana, M.,
Sever, O., Augustynczik, A . L. D., Grote, R., Kyselova, I., Mäkelä, A .,
Yousefpour, R., Krejza, J., Collalti, A., & Reyer, C. P. O. (2019). Forest
carbon allocation modelling under climate change. Tree Physiology,
39, 1937– 1960.
Mesas- Carrascosa, F.- J., de Castro, A. I., Torres- Sánchez, J., Triviño-
Tarradas, P., Jiménez- Brenes, F. M., García- Ferrer, A., & López-
Granados, F. (2020). Classification of 3D point clouds using color
vegetation indices for precision viticulture and digitizing applica-
tions. Remote Sensing, 12(2), 317.
Michel A., Prescher A.- K., Seidling W., & Ferretti M. (2018). ICP forests
brief #1. https://icp- fores ts.org/pdf/ICPFo rests Brief No1.pdf
Mikkelsen, T. N., Clarke, N., Danielewska, A., & Fischer, R. (2013).
Towards supersites in forest ecosystem monitoring and research.
Developments in Environmental Science, 13, 475– 496.
Mills, G., Pleijel, H., Malley, C. S., Sinha, B., Cooper, O. R., Schultz, M.
G., Neufeld, H. S., David, S. D., Sharps, K., Feng, Z., Gerosa, G.,
Harmens, H., Kobayashi, K., Saxena, P., Paoletti, E., Sinha, V., &
Xu, X. (2018). Tropospheric ozone assessment report: Present- day
tropospheric ozone distribution and trends relevant to vegetation.
Elementa: Science of the Anthropocene, 6, 47.
Milović, M., Kebert, M., & Orlović, S. (2021). How mycorrhizas can help
forests to cope with ongoing climate change? Sumarski List, 145(5–
6), 279– 286.
Mitchell, M. J., & Likens, G. E. (2011). Watershed sulfur biogeochemistry:
Shift from atmospheric deposition dominance to climatic regula-
tion. Environmental Science & Technology, 45, 5267– 5271.
Monk s, P. S., Archib ald, A. T., Col ett e, A. , Coope r, O. , Coyle, M., Der went ,
R., Fowler, D., Granier, C., Law, K. S., Mills, G. E., & Stevenson, D.
S. (2015). Tropospheric ozone and its precursors from the urban
to the global scale fromair quality to short- lived climate forcer.
Atmospheric Chemistry and Physics, 15, 8889– 8973.
Moura, B. B., Alves, E. S., Marabesi, M. A., Souza, S. R., Schaub, M., &
Vollenweider, P. (2018). Ozone affects leaf physiology and causes
injury to foliage of native tree species from the tropical Atlantic
Forest of southern Brazil. Science of the Total Environment, 610,
912– 925.
Moura, B. B., Alves, E. S., Souza, S. R., Domingos, M., & Vollenweider,
P. (2014). Ozone phytotoxic potential with regard to fragments of
the Atlantic semi- deciduous Forest downwind of Sao Paulo. Brazil.
Environmental Pollution, 192, 65– 73.
Mozgeris, G., Brukas, V., Pivoriūnas, N., Činga, G., Makrickienė, E.,
Byčenkienė, S., Marozas, V., Mikalajūnas, M., Dudoitis, V., Ulevičius,
   
|
21
DE MARCO et al.
V., & Augustaitis, A. (2019). Spatial pattern of climate change ef-
fects on lithuanian forestry. Forests, 10(9), 809.
Mushinski, R. M., Phillips, R. P., Payne, Z. C., Abney, R. B., Jo, I., Fei,
S., Pusedef, S. E., White, J. R., Rusch, D. B., & Raff, J. D. (2019).
Microbial mechanisms and ecosystem flux estimation for aero-
bic NOy emissions from deciduous forest soils. Proceedings of the
National Academy of Sciences of the United States of America, 116 (6),
2138– 2145.
Myhre, G., Shindell, D., Bréon, F.- M., Collins, W., Fuglestvedt, J., Huang,
J., Koch, D., Lamarque, J.- F., Lee, D., Mendoza, B., Nakajima,
T., Robock, A., Stephens, G., Takemura, T., & Zhang, H. (2013).
Anthropogenic and natural radiative forcing. In T. F. Stocker, D. Qin,
G.- K. Plattner, M. Tignor, S. K. Allen, J. Doschung, A. Nauels, Y. Xia,
V. Bex, & P. M. Midgley (Eds.), Climate change 2013: The physical
science basis. Contribution of working group I to the fifth asse ssment
report of the intergovernmental panel on climate change (pp. 659
740). Cambridge University Press. https://doi.org/10.1017/CBO97
81107 415324.018
Naidoo, S., Slippers, B., Plett, J. M., Coles, D., & Oates, C. N. (2019). The
road to resistance in forest trees. Frontiers In Plant. Science, 29(10),
273. https://doi.org/10.3389/fpls.2019.00273
Nakahara, O., Takahashi, M., Sase, H., Yamada, T., Matsuda, K.,
Ohizumi, T., Fukuhara, H., Inoue, T., Takahashi, A., Kobayashi, H.,
Hatano, R., & Hakamata, T. (2010). Soil and stream water acidifi-
cation in a forested catchment in central Japan. Biogeochemistry,
97, 141– 158.
Niinemets, Ü. (2010). Responses of forest trees to single and multi-
ple environmental stresses from seedlings to mature plants: Past
stress history, stress interactions, tolerance and acclimation. Forest
Ecology and Management, 260, 1623– 1639.
Nilsson, J. (1988). Critical loads for sulphur and nitrogen. In P. Mathy
(Ed.), Air pollution and ecosystems (pp. 85– 91). Springer.
Noce, S., Collalti, A., & Santini, M. (2017). Likelihood of changes in for-
est species suitabilit y, distribution, and diversity under future cli-
mate: The case of Southern Europe. Ecology and Evolution, 7(22),
9358– 9375.
Ohara, T., Akimoto, H., Kurokawa, J., Horii, N., Yamaji, K., Yan, X., &
Hayasaka, T. (2007). An Asian emission inventory of anthropogenic
emission sources for the period 1980– 2020. Atmospheric Chemistry
and Physics, 7, 44194444.
Ohashi, S., Kuroda, K., Takano, T., Suzuki, Y., Fujiwara, T., Abe, H.,
Kagawa, A., Sugiyama, M., Kubojima, Y., Zhang, C., & Yamamoto,
K. (2017). Temporal trends in 137Cs concentrations in the bark, sap-
wood, heartwood, and whole wood of four tree species in Japanese
forests from 2011 to 2016. Journal of Environmental Radioactivity,
178– 179, 335– 342.
Oksanen, E., & Kontunen- Soppela, S. (2021). Plants have different
strategies to defend against air pollutants. Current Opinion in
Environmental Science & Health, 19, 100222.
Paoletti, E., Alivernini, A., Anav, A., Badea, O., Carrari, E., Chivulescu, S.,
Conte, A., Ciriani, M. L., Dalstein- Richier, L., De Marco, A., Fares,
S., Fasano, G., Giovannelli, A., Lazzara, M., Leca, S., Materassi, A.,
Moretti, V., Pitar, D., Popa, I., … Hoshika, Y. (2019). Toward stomatal–
flux based forest protection against ozone: The MOTTLES ap-
proach. Science of the Total Environment, 691, 516– 527.
Paoletti, E., Bytnerowicz, A., Andersen, C., Augustaitis, A., Ferretti, M.,
Grulke, N., Gunthardt- Goerg, M. S., Innes, J., Johnson, D., Karnosky,
D., Luangjame, J., Matyssek, R., McNulty, S., Muller- Starck, G.,
Musselman, R., & Percy, K. (20 07). Impacts of air pollution and cli-
mate change on forest ecosystems— Emerging research needs. The
Scientific World Journal, 7, 18.
Parent, M. B., & Verbyla, D. (2010). The browning of Alaska's boreal for-
est. Remote Sensing, 2(12), 2729– 2747.
Parrent, J. L., & Vilgalys, R. (2007). Biomass and compositional responses
of ec tomycorrhizal fungal hyphae to elevated CO2 and nitrogen fer-
tilization. New Phytologist, 176 , 164174.
Pastorello, G. Z., Trotta, C., Canfora, E., Chu, H., Christianson, D., Cheah,
Y. W., Poindexter, C., Chen, J., Elbashandy, A., Humphrey, M.,
Isaac, P., & Papale, D. (2020). The FLUXNET2015 dataset and the
ONEFlux processing pipeline for eddy covariance data. Scientific
Data, 7, 225. h t t p s : / / d o i . o r g / 1 0 . 1 0 3 8 / s 4 1 5 9 7 - 0 2 0 - 0 5 3 4 - 3
Pecchi, M., Marchi, M., Burton, V., Giannetti, F., Moriondo, M., Bernetti,
I., Bindi, M., & Chirici, G. (2019). Species distribution modelling
to support forest management. A literature review. Ecological
Modelling, 411, 108817.
Peláez, L. M. G., Santos, J. M., Albuquerque, T. T. A., Reis, N. C., Jr.,
Andreão, W. L., & Andrade, M. F. (2020). Air quality status and
trends over large cities in South America. Environmental Science &
Policy, 114, 422– 435.
Pellegrini, A. F. A., Refsland, T., Averill, C., Terrer, C., Staver, A. C.,
Brockway, D. G., Caprio, A., Clatterbuck, W., Coetsee, C., Haywood,
J. D., Hobbie, S. E., Hoffmann, W. A., Kush, J., Lewis, T., Moser,
W. K., Overby, S. T., Patterson, W. A., Peay, K. G., Reich, P. B.,
Jackson, R. B. (2021). Decadal changes in fire frequencies shift tree
communities and functional traits. Nature Ecology & Evolution, 5,
504– 512.
Peng, X., Sonne, C., Lam, S. S., Sik Ok, Y., & Alstrup, A. K. O. (2020).
The ongoing cut- down of the Amazon rainforest threatens the
climate and requires global tree planting projects: A short review.
Environmental Research, 181, 108887.
Peñuelas, J., Poulter, B., Sardans, J., Ciais, P., van der Velde, M., Bopp, L.,
Boucher, O., Godderis, Y., Hinsinger, P., Llusia, J., & Nardin, E. (2013).
Human- induced nitrogen- phosphorus imbalances alter natural and
managed ecosystems across the globe. Nature Communications,
4(1), 2934.
Peñuelas, J., & Staudt, M. (2010). BVOCs and global change. Trends in
Plant Science, 15, 133– 144.
Perera, A. H., Sturtevant, B. R., & Buse, L . J. (2015). Simulation modeling
of forest landscape disturbances. Springer International Publishing.
Perino, A., Pereira, H. M., Navarro, L. M., Fernández, N., Bullock, J.
M., Ceauşu, S., Cortés- Avizanda, A., van Klink, R., Kuemmerle, T.,
Lomba, A., Pe'er, G., & Wheeler, H. C. (2019). Rewilding complex
ecosystems. Science, 364(6438), eaav5570.
Pliūra, A., Jankauskienė, J., Bajerkevičienė, G., Lygis, V., Suchockas,
V., Labokas, J., & Verbylaitė, R. (2019). Response of juveniles of
seven forest tree species and their populations to different com-
binations of simulated climate change- related stressors: Spring-
frost, heat, drought, increased UV radiation and ozone concen-
tration under elevated CO2 level. Journal of Plant Research, 132,
789– 811.
Pope, R. J., Arnold, S. R ., Chipperfield, M. P., Reddington, C. L. S., Butt,
E. W., Keslake, T. D., Feng, W., Latter, B. G., Kerridge, B. J., Siddans,
R., Rizzo, L., Artaxo, P., Sadiq, M., & Tai, A. P. K. (2019). Substantial
increases in Eastern Amazon and Cerrado biomass burning- sourced
tropospheric ozone. Geophysical Research Letters, 46, 1– 22.
Porté, A., & Bartelink, H. H. (2002). Modelling mixed forest growth: A
review of models for forest management. Ecological Modelling, 150,
1 4 1 1 8 8 .
Pourret, O., & Bollinger, J. C. (2018). “Heavy metal”— What to do now:
To use or not to use? Science of the Total Environment, 610– 611,
419– 42 0.
Pulford, I. D., & Watson, C. (2003). Phytoremediation of heavy metal-
contaminated land by trees— A review. Environment International,
29, 529540.
Qiao, Y., Feng, J., Liu, X., Wang, W., Zhang, P., & Zhu, L. (2016). Surface
water pH variations and trends in China from 2004 to 2014.
Environmental Monitoring and Assessment, 188, 443.
Qiu, Y., Guo, L., Xu, X., Zhang, L., Zhang, K., Chen, M., Zhao, Y., Burkey,
K. O., Shew, H. D., Zobel, R. W., Zhang, Y., & Hu, S. (2021). Warming
and elevated ozone induce tradeoffs between fine roots and
mycorrhizal fungi and stimulate organic carbon decomposition.
Science Advances, 7(28), 1– 10.
22 
|
    DE MARCO et al.
Reis, C. R. G., Pacheco, F. S., Reed, S. C., Tejada, G., Nardoto, G. B., Forti,
M. C., & Ometto, J. P. (2020). Biological nitrogen fixation across
major biomes in Latin America: Patterns and global change effects.
Science of The Total Environment, 746, 140998.
Rinnan, R., & Albers, C. N. (2020). Soil uptake of volatile organic com-
pounds: Ubiquitous and underestimated? JGR Biogeosciences, 125,
e2020JG005773.
Rizzetto, S., Belyazid, S., Gégout, J.- C., Nicolas, M., Alard, D., Corcket, E.,
Gaudio, N., Sverdrup, H., & Probst, A. (2016). Modelling the impact
of climate change and atmospheric N deposition on French forests
biodiversity. Environmental Pollution, 213, 1016– 1027.
Rockström, J., Steffen, W., Noone, K., Persson, Å., Chapin, F. S., III,
Lambin, E. F., Lenton, T. M., Scheffer, M., Folke, C., Schellnhuber, H.
J., Nyk vist, B., de Wit, C. A., Hughes, T., van der Leeuw, S., Rodhe,
H., Sörlin, S., Snyder, P. K., Costanza, R., Svedin, U., Foley, J. A.
(2009). A safe operating space for humanity. Nature, 461, 472– 475.
Rodman, K. C., Andrus, R. A., Veblen, T. T., & Hart, S. J. (2021).
Disturbance detection in landsat time series is influenced by tree
mortality agent and severity, not by prior disturbance. Remote
Sensing of Environment, 254, 112244.
Rogers, B. M., Solvik, K., Hogg, E. H., Ju, J., Masek, J. G., Michaelian,
M., Berner, L. T., & Goetz, S. J. (2018). Detecting early warning
signals of tree mortality in boreal North America using multiscale
satellite data. Global Change Biology, 24(6), 2284– 2304. https://d o i .
org /10.1111/gcb.14107
Saitanis, C. J., Agathokleous, E., Burkey, K ., & Hung, Y. T. (2020). Chapter
8. Ground level ozone profile and the role of plants as sources and
sinks. In Y. T. Hung, L. K. Wang, & N. Shammas (Eds.), Handbook of en-
vironment and waste management, Vol. 3: Acid rain and greenhouse gas
pollution control (pp. 281324). World Scientific Publishing Co. Inc.
Saitanis, C. J., Sicard, P., De Marco, A., Feng, Z., Paoletti, E., & Agathokleous,
E. (2020). On the atmospheric ozone monitoring methodologies.
Current Opinion in Environmental Science & Health, 18, 40– 46.
Salojär vi, J., Smolander, O.- P., Nieminen, K., Rajaraman, S., Safronov, O.,
Safdari, P., Lamminmäki, A., Immanen, J., Lan, T., Tanskanen, J., &
Rastas, P. (2017). Genome sequencing and population genomic
analyses provide insights into the adaptive landscape of silver birch.
Nature Genetics, 49, 904– 912.
Sandermann, H., Wellburn, A . R., & Heath, R. L. (1997). Forest decline and
ozone: Synopsis. Springer.
Santini, M., Collalti, A., & Valentini, R. (2014). Climate change impacts
on vegetation and water cycle in the Euro- Mediterranean region,
studied by a likelihood approach. Regional Environmental Change,
14(4), 1405– 1418.
Sardans, J., Alonso, R., Carnicer, J., Fernández- Martínez, M., Vivanco, M.
G., & Peñuelas, J. (2016). Factors influencing the foliar elemental
composition and stoichiometry in forest trees in Spain. Perspectives
in Plant Ecology, Evolution and Systematics, 18, 52– 69.
Sase, H., Saito, T., Takahashi, M., Morohashi, M., Yamashita, N., Inomata,
Y., Ohizumi, T., & Nakata, M. (2021). Transboundary air pollution
reduction rapidly reflected in stream water chemistry in forested
catchment on the Sea of Japan coast in central Japan. Atmospheric
Environment, 248, 118223.
Sase, H., Takahashi, M., Matsuda, K., Sato, K., Tanikawa, T., Yamashita,
N., Ohizumi, T., Ishida, T., Kamisako, M., Kobayashi, R., Uchiyama,
S., Saito, T., Morohashi, M., Fukuhara, H., Kaneko, S., Inoue, T.,
Yamada, T., Takenaka, C., Tayasu, I., … Ohta, S. (2019). Response
of river water chemistry to changing atmospheric environment
and sulfur dynamics in a forested catchment in central Japan.
Biogeochemistry, 142, 357– 374.
Sase, H., Yamashita, N., Luangjame, J., Garivait, H., Kietvuttinon, B.,
Visaratana, T., Kamisako, M., Kobayashi, R., Ohta, S., Shindo, J.,
Hayashi, K., Toda, H., & Matsuda, K. (2017). Alkalinization and
acidification of stream water with changes in atmospheric depo-
sition in a tropical dry evergreen forest of northeastern Thailand.
Hydrological Processes, 31, 836– 846.
Schaub, M., Häni, M., Calatayud, V., Ferretti, M., & Gottardini, E. (2020).
Ozone concentrations are decreasing but exposure remains high in
European forests ICP Forest s Brief. ICP Forests.
Schindler, T., Mander, Ü., Machacova, K., Espenberg, M., Krasnov, D.,
Escuer- Gatius, J., Veber, G., Pärn, J., & Soosaar, K. (2020). Short-
term flooding increases CH4 and N2O emissions from trees in a ri-
parian forest soil- stem continuum. Scientific Reports, 10, 3204.
Schlutow, A., Schröder, W., & Scheuschner, T. (2021). Assessing the
relevance of atmospheric heavy metal deposition with regard to
ecosystem integrity and human health in Germany. Environmental
Science Eurrope, 33, 7.
Schmitz, A., Sanders, T. G. M., Bolte, A., Bussotti, F., Dirnböck, T.,
Johnson, J., Peñuelas, J., Pollastrini, M., Prescher, A.- K., Sardans,
J., Verstraeten, A., & de Vries, W. (2019). Responses of forest eco-
systems in Europe to decreasing nitrogen deposition. Environmental
Pollution, 244, 980994.
Schultz, M. G., Schröder, S., Lyapina, O., Cooper, O. R., Galbally, I.,
Petropavlovskikh, I., Von Schneidemesser, E., Tanimoto, H., Elshorbany,
Y., Naja, M., Seguel, R. J., Dauert, U., Eckhardt, P., Feigenspan, S., Fiebig,
M., Hjellbrekke, A. G., Hong, Y. D., Kjeld, P. C., Koide, H., … Zhiqiang, M.
(2017). Tropospheric Ozone Assessment Report: Database and metrics
data of global surface ozone observations. Elementa, 5, 58.
Schwede, D. B., Simpson, D., Tan, J., Fu, J. S., Dentener, F., Du, E., & Vries,
W. (2018). Spatial variation of modelled total, dry and wet nitrogen
deposition to forests at global scale. Environmental Pollution, 243,
1287– 1301.
Serengil, Y., Augustaitis, A., Bytnerowicz, A., Grulke, N., Kozovitz, A.
R., Matyssek, R., Müller- Starck, G., Schaub, M., Wieser, G., Aydin
Coskun, A., & Paoletti, E. (2011). Adaptation of forest ecosystems
to air pollution and climate change: A global assessment on re-
search priorities. iForest- Biogeosciences and Forestry, 4, 44– 48.
Shestakov, A. L., Filippov, B. Y., Zubrii, N. A., Klemola, T., Zezin, I., Zverev,
V., Zvereva, E. L., & Kozlov, M. V. (2020). Doubling of biomass pro-
duction in European boreal forest trees by a four- year suppression
of background insect herbivory. Forest Ecology and Management,
462, 117992. https://doi.org/10.1016/j.foreco.2020.117992
Shi, C., Watanabe, T., & Koike, T. (2017). Leaf stoichiometry of deciduous tree
species in different soils exposed to free- air O3 enrichment over two
growing seasons. Environmental and Experimental Botany, 138, 148– 163.
Shifley, S. R., He, H. S., Lischke, H., Wang, W. J., Jin, W., Gustafson, E. J.,
Thompson, J. R., Thompson, F. R., Dijak, W. D., & Yang, J. (2017). The
past and future of modeling forest dynamics: From growth and yield
curves to forest landscape models. Landscape Ecology, 32, 1307– 1325.
Sicard, P. (2021). Ground- level ozone over time: An observation- based
global overview. Current Opinion in Environmental Science & Health,
19, 100226.
Sicard, P., Agathokleous, E., Araminiene, V., Carrari, E., Hoshika, Y.,
De Marco, A., & Paoletti, E. (2018). Should we see urban trees
as effective solutions to reduce increasing ozone levels in cities?
Environmental Pollution, 243A , 163– 176.
Sicard, P., Anav, A., De Marco, A., & Paoletti, E. (2017). Projected global
ground- level ozone impacts on vegetation under different emis-
sion and climate scenarios. Atmospheric Chemistry and Physics, 17,
12177– 12196.
Sicard, P., Augustaitis, A., Belyazid, S., Calfapietra, C., De Marco, A., Fenn,
M., Bytnerowicz, A., Grulke, N., He, S., Matyssek, R., & Serengil, Y.
(2016). Global topics and novel app ro ac he s in the study of air poll u-
tion, climate change and forest ecosystems. Environmental Pollution,
213, 977– 987.
Sicard, P., De Marco, A., Carrari, E., Dalstein- Richier, L., Hoshika, Y.,
Badea, O., Pitar, D., Fares, S., Conte, A., Popa, I., & Paoletti, E.
(2020). Epidemiological derivation of flux- based critical levels
for visible ozone injury in European forests. Journal of Forestry
Research, 31, 1509– 1519.
Sicard, P., De Marco, A., Carrari, E., Hoshika, Y., & Paoletti, E. (2021).
Testing visible ozone injury within a Light Exposed Sampling Site as
   
|
23
DE MARCO et al.
a proxy for ozone risk assessment for European forests. Journal of
Forestry Research, 32, 1351– 1359.
Sicard, P., De Marco, A., Dalstein- Richier, L., Tagliaferro, F., Renou, C.,
& Paoletti, E. (2016). An epidemiological assessment of stomatal
ozone flux- based critical levels for visible ozone injur y in southern
European forests. Science of the Total Environment, 5 41, 729– 741.
Silfver, T., Heiskanen, L., Aurela, M., Myller, K., Karhu, K., Meyer, N.,
Oksanen, E., Rousi, M., & Mikola, J. (2020). Insect herbivory con-
trol of Subarctic ecosystem CO2 exchange in present and future
climates. Nature Communications, 11, 2529.
Simard, S. (Ed.). (2010). Climate change and variability. IntechOpen.
htt ps://doi.org /10.5772/1743
Šimpraga, M., Ghimire, R. P., Van Der Straeten, D., Blande, J. D., Kasurinen,
A., Sorvari, J., Holopainen, T., Adriaenssens, S., Holopainen, J. K., &
Kivimäenpää, M. (2019). Unravelling the functions of biogenic vol-
atiles in boreal and temperate forest ecosystems. European Journal
of Forest Research, 138, 763– 787.
Sniezko, R. A ., & Koch, J. (2017). Breeding trees resistant to insects and
diseases: Putting theory into application. Biological Invasions, 19,
33 7 7– 3 4 0 0 .
Souza, M. A., Pacheco, F. S., Palandi, J. A. L., Forti, M. C., Campos, L. A.
M., Ometto, J. P. H. B., Reis, D. C. O., & Carvalho Junior, J. A. (2020).
Atmospheric concentrations and dry deposition of reactive nitrogen in
the state of São Paulo, Brazil. Atmospheric Environment, 230, 117502.
Stankevich, S. A., Kozlova, A. A., Piestova, I. O., & Lubskyi, M. S. (2017).
Leaf area index estimation of forest using sentinel- 1 C- band SAR
data. In 2017 IEEE microwaves, Radar and Remote Sensing Symposium
(MRRS) (pp. 253– 256). IEEE.
Steffen, W., Richardson, K., Rockström, J., Cornell, S., Fetzer, I., Bennett,
E., Biggs, R., Carpenter, S. R., de Wit, C., Folke, C., Mace, G. M.,
Persson, L. M., Veerabhadran, R., Reyers, B., & Sörlin, S. (2015).
Planetary Boundaries: Guiding human development on a changing
planet. Science, 347, 1259855.
Stoddard, J. L., Jeffries, D. S., Lükewille, A., Clair, T. A., Dillon, P. J.,
Driscoll, C. T., Forsius, M., Johannessen, M., Kahl, J. S., Kellogg, J.
H., Kemp, A., Mannio, J., Monteith, D. T., Murdoch, P. S., Patrick, S.,
Rebsdorf, A., Skjelkvåle, B. L., Stainton, M. P., Traaen, T., … Wilander,
A. (1999). Regional trends in aquatic recovery from acidification in
North America and Europe. Nature, 401, 575– 578.
Strand, P., Sundell- Bergman, S., Brown, J. E., & Dowdall, M. (2017). On
the divergences in assessment of environmental impacts from
ionising radiation following the Fukushima accident. Journal of
Environmental Radioactivity, 169– 1 70, 159– 173.
Suchara, I., Sucharova, J., Hola, M., Pilatova, H., & Rulik, P. (2016). Long-
term retention of 137Cs in three forest soils with dif ferent soil prop-
erties. Journal of Environmental Radioactivity, 158– 159, 102113.
Sugai, T., Yannan, W., Watanabe, T., Satoh, F., Qu, L., & Koike, T. (2019).
Salt stress reduced the seedling growth of two larch species under
elevated ozone. Frontiers in Fores ts and Global Change, 2, 53.
Sverdrup, H., & De Vries, W. (1994). Calculating critical loads for acidity
with the simple mass balance method. Water, Air, & Soil Pollution,
72, 143– 162.
Takahashi, M., Feng, Z., Mikhailova, T. A., Kalugina, O. V., Shergina, O.
V., Larisa, V., Afanasieva, K. J., Heng, R., Majid, N. M. A., & Sase, H.
(2020). Air pollution monitoring and tree and forest decline in East
Asia: A review. Science of the Total Environment, 742 , 140288.
Talhelm, A. F., Pregitzer, K. S., Kubiske, M. E., Zak, D. R., Campany, C. E.,
Burton, A. J., Dickson, R. E., Hendrey, G. R., Isebrands, J. G., Lewin,
K. F., Nagy, J., & Karnosky, D. F. (2014). Elevated carbon dioxide and
ozone alter productivity and ecosystem carbon content in northern
temperate forests. Global Change Biology, 20, 2492– 250 4. h t t p s : //d o i .
org /10.1111/gcb.12 56 4
Tamaoki, M. (2016). Studies on radiation effects from the Fukushima
nuclear accident on wild organisms and ecosystems. Global
Environmental Research, 20, 73– 82.
Tănase, M. A., Villard, L., Pitar, D., Apostol, B., Petrila, M., Chivulescu, S.,
Leca, S., Borlaf- Mena, I., Pascu, I. S., Dobre, A. C., Pitar, D., Guiman,
G., Lorent, A., Anghelus, C., Ciceu, A ., Nedea, G., Stanculeanu, R .,
Popescu, F., Aponte, C., & Badea, O. (2019). SYNTHETIC APERTURE
RADAR sensitivity to forest changes: A simulations- based study
for the Romanian forests. Science of the Total Environment, 689,
1104– 1114.
Tani, A., & Mochizuki, T. (2021). Review: Exchanges of volatile organic
compounds between terrestrial ecosystems and the atmosphere.
Journal of Agricultural Meteorology, 77, 66– 80.
Tenkanen, A., Keski- Saari, S., Salojärvi, J., Oksanen, E., Keinänen, M., &
Kontunen- Soppela, S. (2020). Differences in growth and gas ex-
change between southern and northern provenances of silver birch
(Betula pendula) in northern Europe. Tree Physiology, 40, 198– 214.
Terada, H., Nagai, H., Tsuduki, K., Furuno, A., Kadowaki, M., & Kakefuda,
T. (2020). Refinement of source term and atmospheric dispersion
simulations of radionuclides during the Fukushima Daiichi Nuclear
Power Station accident. Journal of Environmental Radioactivity, 213,
106104.
Terrer, C., Jackson, R. B., Prentice, I. C., Keenan, T. F., Kaiser, C., Vicca, S.,
Fisher, J. B., Reich, P. B., Stocker, B. D., Hungate, B. A., Penuelas, J.,
McCallum, I., Soudzilovskaia, N. A., Cernusak, L. A., Talhelm, A. F.,
Sundert, K. V., Piao, S., Newton, P. C. D., Hovenden, M. J., … Franklin,
O. (2019). Nitrogen and phosphorus constrain the CO2 fertilization of
global plant biomass. Nature Climate Change, 9, 684– 689.
Tian, D., Du, E., Jiang, L., Ma, S., Zeng, W., Zou, A., Feng, C., Xu, L., Xing,
A., Wang, W., & Zheng, C. (2018). Responses of forest ecosystems
to increasing N deposition in China: A critical review. Environmental
Pollution, 243, 75– 86.
Torres, P., Rodes- Blanco, M., Viana- Soto, A., Nieto, H., & García, M. (2021).
The role of remote sensing for the assessment and monitoring of for-
est health: A systematic evidence synthesis. Forests, 12, 1134.
Tørseth, K., Aas, W., Breivik, K., Fjæraa, A. M., Fiebig, M., Hjellbrekke,
A. G., Lund, M. C., Solberg, S., & Yttri, K. E. (2012). Introduction
to the European Monitoring and Evaluation Programme (EMEP)
and observed atmospheric composition change during 1972– 2009.
Atmospheric Chemistry and Physics, 12, 5447– 5481.
Tóth, G., Hermann, T., Szatmári, G., & Pásztor, L. (2016). Maps of heavy
metals in the soils of the European Union and proposed priority
areas for detailed assessment. Science of the Total Environment, 565,
1054– 1062.
Tracy, S. R., Lilleskov, N. K. A., Postma, J. A., Fassbender, H., Wasson,
A., & Watt, M. (2020). Crop improvement from phenotyping roots:
Highlights reveal expanding opportunities. Trends in Plant Science,
25(1), 105– 118.
United Nations (Ed.). (2000). Sources and effects of ionizing radia-
tion: United Nations Scientific Committee on the Effects of Atomic
Radiation: UNSCEAR 2000 report to the General Assembly, with scien-
tific annexes. United Nations.
United States Federal Register. (2015). Federal Register, Vol. 80, No. 225,
Monday, November 23, 2015, Notices.
Uppala, S. M., Kallberg, P. W., Simmons, A. J., Andrae, U., Da Costa
Bechtold, V., Fiorino, M., Gibson, J. K., Haseler, J., Hernandez,
A., Kelly, G. A., & Li, X. (2005). The ERA- 40 re- analysis. Quarterly
Journal of the Royal Meteorological Societ y, 131(612), 2961– 3 012.
Urban, M. C. (2015). Accelerating extinction risk from climate change.
Science, 348(6234), 571– 573.
Urban, M. C., Bocedi, G., Hendry, A. P. P., Mihoub, J.- B., Pe'er, G., Singer,
A., Bridle, J. R., Crozier, L. G., de Meester, L., Godsoe, W., Gonzalez,
A., Hellmann, J. J., Holt, R. D., Huth, A., Johst, K., Krug, C. B.,
Leadley, P. W., Palmer, S. C. F., Pantel, J. H., … Travis, J. M. J. (2016).
Improving the forecast for biodiversity under climate change.
Science, 353(6304), aad8466.
van der Linde, S., Suz, L. M., Orme, C., David, L., Cox, F., Andreae, H.,
Asi, E., Atkinson, B., Benham, S., Carroll, C., Cools, N., & De Vos, B.
24 
|
    DE MARCO et al.
(2018). Environment and host as large- scale controls of ectomycor-
rhizal fungi. Nature, 558 , 243248.
Varghese, A., Ticktin, T., Mandle, L., & Nath, S. (2015). Assessing the
effects of multiple stressors on the recruitment of fruit harvested
trees in a Tropical Dry Forest, Western Ghats, India. PLoS One,
10(3), e0119634.
Vuorenmaa, J., Augustaitis, A., Beudert, B., Clarke, N., de Wit, H. A.,
Dirnböck, T., Frey, J., Forsius, M., Indriksone, I., Kleemolaa, S., Kobler,
J., Krám, P., Lindroos, A.- J., Lundin, L., Ruoho- Airola, T., Ukonmaanaho,
L., & Váňa, M. (2017). Long- term sulphate and inorganic nitrogen mass
balance budgets in European ICP Integrated Monitoring catchments
( 1 9 9 0 2 0 1 2 ) . Ecological Indicators, 76, 15– 29.
Wamelink, G. W. W., Mol- Dijkstra, J. P., Reinds, G. J., Voogd, J. C.,
Bonten, L. T. C., Posch, M., Hennekens, S. M., & De Vries, W. (2020).
Prediction of plant species occurrence as affected by nitrogen
deposition and climate change on a European scale. Environmental
Pollution, 266(2), 115257.
Wang, X., Qu, L., Mao, Q., Watanabe, M., Hoshika, Y., Koyama, A.,
Kawaguchi, K ., Tamai, Y., & Koike, T. (2015). Ectomycorrhizal colo-
nization and growth of the hybrid larch F1 under elevated CO2 and
O3. Environmental Pollution, 197, 116– 126.
Watanabe, M., Hoshika, Y., Koike, T., & Izuta, T. (2017). Combined effects
of ozone and other environmental factors on Japanese trees. In T.
Izuta (Ed.), Air pollution impacts on plant in East Asia (pp. 101110).
Springer.
Watanabe, Y., Ichikawa, S., Kubota, M., Hoshino, J., Kubota, Y., Maruyama,
K., Fuma, S., Kawaguchi, I., Yoschenko, V. I., & Yoshida, S. (2015).
Morphological defects in native Japanese fir trees around the
Fukushima Daiichi Nuclear Power Plant. Scientific Reports, 5, 13232.
Wei, W., Cheng, S., Li, G., Wang, G., & Wang, H. (2014). Characteristics
of ozone and ozone precursors (VOCs and NOx) around a petro-
leum refinery in Beijing, China. Journal of Environmental Sciences,
26, 332– 342.
Wenig, M., Ghirardo, A., Sales, J. H., Pabst, E. S., Breitenbach, H. H.,
Antritter, F., Weber, B., Lange, B., Lenk, M., Cameron, R. K., &
Schnitzler, J. P. (2019). Systemic acquired resistance networks am-
plify airborne defense cues. Nature Communications, 10, 3813.
Wieder, W. R., Cleveland, C. C., Smith, W. K., & Todd- Brown, K. (2015).
Future productivity and carbon storage limited by terrestrial nutri-
ent availability. Nature Geoscience, 8(6), 441444.
Wiley, E., King, C. M., & Landhäusser, S. M. (2020). Identifying the rel-
evant carbohydrate storage pools available for remobilization in
aspen roots. Tree Physiology, 39, 1109– 1120.
WMO, GAW. (2003). Aerosol measurement procedures, guidelines and rec-
ommendations. GAW report.
Xie, D., Si, G., Zhang, T., Mulder, J., & Duan, L. (2018). Nitrogen deposi-
tion inc rease s N2O emission from an N- saturated subtropical forest
in southwest China. Environmental Pollution, 243, 1818– 1824.
Yao, Z., Zheng, X., Xie, B., Liu, C., Mei, B., Dong, H., Butterbach- Bahl, K.,
& Zhu, J. (2009). Comparison of manual and automated chambers
for field measurements of N2O, CH4, CO2 fluxes from cultivated
land. Atmospheric Environment, 43(11), 1888– 1896.
Yoschenko, V., Nanba, K., Yoshida, S., Watanabe, Y., Takase, T., Sato, N.,
& Keitoku, K. (2016). Morphological abnormalities in Japanese red
pine (Pinus densiflora) at the territories contaminated as a result of
the accident at Fukushima Dai- Ichi Nuclear Power Plant. Journal of
Environmental Radioactivity, 165, 60– 67.
Yoschenko, V. I., Kashparov, V. A., Melnychuk, M. D., Levchuk, S. E.,
Bondar, Y. O., Lazarev, M., Yoschenko, M. I., Farfán, E. B., & Jannik,
G. T. (2011). Chronic irradiation of Scots pine trees (Pinus sylvestris)
in the Chernobyl Exclusion Zone: Dosimetry and radiobiological ef-
fects. Health Physics, 101, 393– 408.
Yoshida, N., & Takahashi, Y. (2012). Land- surface contamination by ra-
dionuclides from the Fukushima Daiichi nuclear power plant acci-
dent. Elements, 8(3), 201– 206.
Yu, H., & Blande, J. D. (2021). Diurnal variation in BVOC emission and
CO2 gas exchange from above- and belowground parts of two co-
niferous species and their responses to elevated O3. Environmental
Pollution, 278, 116830.
Yue, C., Cui, K., Duan, J., Wu, X., Rodriguez, C., Fu, H., Deng, T., Zhang,
S., Liu , J., Guo, Z., Xi, B., & Cao, Z. (2021). The retention characteris-
tics for water- soluble and water- insoluble particulate matter of five
tree species along an air pollution gradient in Beijing, China. Science
of the Total Environment, 767, 145497.
Yue, X., & Unger, N. (2018). Fire air pollution reduces global terrestrial
productivity. Nature Communications, 9, 5413.
Zaehle, S., Friedlingstein, P., & Friend, A. D. (2010). Terrestrial nitro-
gen feedbacks may accelerate future climate change. Geophysical
Research Letters, 37, L01401.
Zang, M. G., Zhou, Z. K., Chen, W. Y., Slik, J. W. F., Cannon, C. H., &
Raes, N. (2012). Using species distribution modeling to improve
conservation and land use planning of Yunnan, China. Biological
Conservation, 153, 257– 264.
Zhang, M., Wang, W., Tang, L., Heenan, M., Wang, D., & Xu, Z. (2021).
Impacts of prescribed burning on urban forest soil: Minor changes
in net greenhouse gas emissions despite evident alterations of mi-
crobial community structures. Applied Soil Ecology, 158, 103780.
Zheng, B., Tong, D., Li, M., Liu, F., Hong, C., Geng, G., Li, H., Li, X., Peng,
L., Qi, J., Yan, L., Zhang, Y., Zhao, H., Zheng, Y., He, K., & Zhang,
Q. (2018). Trends in China's anthropogenic emissions since 2010
as the consequence of clean air actions. Atmospheric Chemistry and
Physics, 18, 14095– 14111.
Zhong, Q., Shen, H., Yun, X., Chen, Y., Ren, Y., Xu, H., Shen, G., Du, W.,
Meng, J., Li, W., Ma, J., & Tao, S. (2020). Global sulfur dioxide emis-
sions and the driving forces. Environmental Science & Technology, 54,
6508– 6517.
Zvereva, E. L., & Kozlov, M. V. (2006). Consequences of simultaneous
elevation of carbon dioxide and temperature for plant- herbivore
interactions: A meta- analysis. Global Change Biology, 12, 27– 41.
Zvereva, E. L., & Kozlov, M. V. (2010). Responses of terrestrial arthropods
to air pollution: A meta- analysis. Environmental Science and Pollution
Research, 17, 297– 311.
Zvereva, E. L., Roitto, M., & Kozlov, M. V. (2010). Growth and reproduc-
tion of vascular plants under pollution impact: A synthesis of exist-
ing knowledge. Environmental Reviews, 18, 355– 367.
Zvereva, E. L., Toivonen, E., & Kozlov, M. V. (2008). Changes in species
richness of vascular plants under pollution impact: A global per-
spective. Global Ecology and Biogeography, 17, 305– 319.
Zvereva, E. L., Zverev, V. E., & Kozlov, M. V. (2012). Little strokes fell
great oaks: Minor but chronic herbivory substantially reduces birch
growth. Oikos, 121, 2036– 2043.
SUPPORTING INFORMATION
Additional supporting information may be found in the online
version of the article at the publisher’s website.
How to cite this article: De Marco, A., Sicard, P., Feng, Z.,
Agathokleous, E., Alonso, R., Araminiene, V., Augustatis, A.,
Badea, O., Beasley, J. C., Branquinho, C., Bruckman, V. J.,
Collalti, A., David- Schwartz, R., Domingos, M., Du, E., Garcia
Gomez, H., Hashimoto, S., Hoshika, Y., Jakovljevic, T. …
Paoletti, E. (2022). Strategic roadmap to assess forest
vulnerability under air pollution and climate change. Global
Change Biology, 00, 124. https://doi.org/10.1111/gcb.16278
... Climate change and air pollution have emerged as one of the most pressing global challenges affecting human health, ecosystems, and biodiversity [1,2]. Awareness has intensified and rapidly increased worldwide [3,4], with the intensification of extreme weather events and rising global temperatures signaling the urgent need for effective mitigation strategies, primarily through the reduction of greenhouse gases [5,6]. ...
... Direct observations determined the species of the trees, the DBH was measured with a tree caliper, and the height was measured with the help of a hypsometer. These measurements were then applied to the bifactorial volume regression Equation (1). The regression coefficients for this equation have been determined for 43 forest species, as described by Giurgiu [52]. ...
... Equation (1). The regression coefficients for this equation have been determined for 43 forest species, as described by Giurgiu [52]. ...
Article
Full-text available
With the intensification of the effects of climate change, the urgent need to address their drivers, especially greenhouse gas emissions, has become essential. In this context, forests offer a robust solution, with their potential to store and mitigate carbon emissions. However, striking a balance is critical given the significant economic contribution of the forestry and wood-based industries, which account for about 5% of Romania’s GDP and employ 6% (around 300 thousand) of its active workforce. This study, conducted in the Piatra Craiului National Park located in Romania’s Southern Carpathians, we utilize the EFISCEN application to generate three distinct 50-year forest evolution scenarios based on harvest intensity, namely Business As Usual (BAU), Maximum Intensity (MAX), and No Harvest (MIN), on two historical different managed forests, i.e., conservation and production. The study aims to guide forest owners in decision making with scenario modeling tools, with the objectives of assessing the forest carbon sequestration potential and evaluating the economic feasibility. In the most probable scenario, the BAU scenario, the growing stock increases from 2.6 million m3 to 3.8 million m3 over 50 years, with a more than 40% increase. Comparing the carbon stock change for all tree harvest scenario types indicates that the MIN scenario has the highest carbon sink capacity in the next 50 years; the BAU scenario is a well-balanced option between carbon sink and wood provision and has an optimal EUR 3.7 million in annual revenue. The MAX scenario can boost the growth and increase the annual revenue from wood by 35% but is effective only for a short time and thus has the smallest calculated revenue in time. Achieving a win–win relationship between carbon sequestration and wood supply is imperative, as well as good planning and scenarios to contribute to climate mitigation and also as provisions for local communities and to sustain the local economy.
... In broader terms, climate change affects forest carbon balance by influencing key processes, 391 which can respond differently due to their sensitivity to various environmental drivers [80]. 392 Indeed, the reduction in NEE, within the forests observed in this study, has a disparity in 393 . ...
Preprint
Full-text available
Through photosynthesis, forests absorb annually large amounts of atmospheric CO2. However, they also release CO2 back through respiration. These two, opposite in sign, large fluxes determine, much of the carbon that is stored or released back to the atmosphere. The mean seasonal cycle (MSC) is an interesting metric that associate phenology and carbon (C) partition-ing-allocation analysis within forest stands. Here we applied the 3D-CMCC-FEM model and analyzed its capability to represent the main C-fluxes, by validating the model against observed data, questioning if the sink/source mean seasonality is influenced under two scenarios of climate change, in five contrasting European forest sites. We found the model has, under current climate conditions, robust predictive abilities in estimating NEE. Model results also predict a consistent reduction of the forest's capabilities to act as a C-sink under climate change and stand-ageing at all sites. Such a reduction is predicted despite the number of annual days of C-sink in evergreen forests increasing over the years, indicating a consistent downward trend. Similarly, deciduous forests, despite maintaining a relatively stable number of C-sink days throughout the year and over the century, show a reduction in their overall annual C-sink capacity. Overall, both types of forests at all sites show a consistent reduction in their future mitigating potential.
... L'attitudine del bosco a riprendersi da stress e disturbi, senza perdere l'integrità strutturale e funzionale, definisce la sua resilienza. In un quadro gestionale, la resilienza può essere considerata come la capacità del sistema di garantire i servizi ecosistemici (Morán-Ordóñez 2020, Lecina-Diaz et al. 2021, De Marco et al. 2022. ...
Article
Full-text available
This study provides an overview of the forests of Basilicata, Southern Italy, including their recent history, dominant forest types, current management, and vulnerability to climate change and wildfire. It outlines silvicultural and management proposals that can be implemented in the new forest plan that the Basilicata Region is about to adopt. The proposals are based on the principle of adaptive management to support forest functionality, biodiversity and ecosystem services. Silvicultural methods include continuous cover forestry, natural regeneration, species richness and functional diversity, structural diversification, imitation of natural disturbances, tree retention to increase biodiversity. The characteristics of the forest-wood supply chain have been analyzed, highlighting weaknesses and possible improvements.
... Fungi Carthamus oxycantha L. ↑Shoot length, root length, total chlorophyll and carotenoids [112] wheat seed endophytic bacteria (WSEB) ...
Article
Full-text available
Endophytes, as microorganisms widely present in plants, have an important role in plant growth and development. Abiotic stresses are very essential influence on plant growth and development. Endophytes in host plants are diverse, however, beneficial endophytes are used to make plants resistant to abiotic stresses. This review focuses on studying the regulatory roles of different endophytes under abiotic stresses, and explained the special pathway and related mechanism of endophytes under heavy metal stress, such as cadmium, manganese and zinc stress. How do the dominant endophytes respond to salt and heat stress and affect plant physiological characteristics? In addition, we also summarized the potential and application of endophytes in reducing the toxicity of plant pathogens, promoting crop growth, biomedicine and ecological restoration, and other aspects, to provide reference for further in-depth research on the mechanism of action of plant endophytes under abiotic stresses and effective utilization of endophytes.
... Global warming causes the water level to rise, which directly affects the sustainable development of humankind [17]. These climate changes affect not only the waters but also the forests; along with industrial pollution and deforestation, global warming leads to devastating forest fires that destroy biodiversity [18]. However, it should also be noted that the human hand can destroy these ecosystems through intentionally setting fires in the forest to facilitate land use change [19]. ...
Article
Full-text available
The concept of sustainable development appeared as a response to the attempt to improve the quality of human life, simultaneously with the preservation of the environment. For this reason, two of the 17 Sustainable Development Goals are dedicated to life below water (SDG14) and on land (SDG15). In the course of this research, comprehensive information on the extent of degradation in Romania’s primary ecosystems was furnished, along with an exploration of the key factors precipitating this phenomenon. This investigation delves into the perspectives of 42 counties, scrutinizing the level of degradation in forest ecosystems, grasslands, lakes and rivers. The analysis commences with a presentation of descriptive statistics pertaining to each scrutinized system, followed by an elucidation of the primary causes contributing to its degradation. Subsequently, a cluster analysis is conducted on the counties of the country. One of these causes is the presence of intense industrial activity in certain areas, so it is even more important to accelerate the transition to a green economy in order to help the environment regenerate.
... Forests have been exposed to extreme events more and more frequently (such as drought spells, heat waves, windstorms, wildfires, and spring frost events). Thus they become less stable due to such climate change (De Marco et al., 2022;Lucas-Borja et al., 2022;Ortega et al., 2023). In fact, these disturbances can change the composition, structure, and functionality of the forests (Cobb, 2022). ...
Article
Forest ecosystems are among the most important carbon sinks in the terrestrial biomes. Therefore, the estimation of aboveground biomass accumulation is of fundamental importance for understanding the contribution of forest stands to the global carbon budget. In this study, we proposed a modelling approach to estimate aboveground biomass of beech forests in the Central Apennine based on stand age, climatic variables, topographic features and soil parameters. Using forest inventory data from the local forest management plans, international databases, and mixed-effect linear model, we identified stand age as a major driver of beech forests aboveground biomass. Climatic variables had generally higher influence than soil and topographic features, probably as a consequence of the homogeneous calcareous substrate which favoured the development of soils which are highly suitable for the growth of beech. Temperature range and seasonality were the most important climatic variables. Interestingly , we found that the aboveground biomass in Apennine beech forests is strongly site-specific. Different management approaches during the past centuries, i.e. presence or absence of conversion interventions of the analysed stands, are probably responsible for the growing site's significant influence. By highlighting the ideal locations for allocating the functions of forest production or evaluating the value of ecosystem services that regulate the climate, this study will help to improve both the precision of carbon budget modelling and decision-making in nature conservation and forest management.
... In Europe alone, forest ecosystems, which cover about a 40%, currently act as a net carbon sink for ~ 315 Megatonnes of CO 2 eq and compensate for about 8% of EU-27's total greenhouse emissions (Verkerk et al., 2022). However, adverse climate impacts such as heat waves and drought (Allen et al., 2015;D'Andrea et al., 2020D'Andrea et al., , 2021Schuldt et al., 2020) and increasing natural disturbance rates (Grünig et al., 2023;Patacca et al., 2023) are all stressors which have potentially significant effects on current and future forest dynamics, jeopardizing the European forest ecosystems functioning and their carbon mitigation potential under future climate change (De Marco et al., 2022;Schuldt et al., 2020;Senf et al., 2020). ...
Article
Full-text available
Process-based Forest Models (PBFMs) offer the possibility to capture important spatial and temporal patterns of both carbon fluxes and stocks in forests, accounting for ecophysiological, climate and geographical variability. Yet, their predictive capacity should be demonstrated not only at the stand-level but also in the context of large spatial and temporal heterogeneity. For the first time, we apply a stand scale process-based model (3D-CMCC-FEM) in a spatially explicit manner at 1 km spatial resolution in a Mediterranean region in southern Italy. Specifically, we developed a methodology to initialize the model that comprehends the use of spatial information derived from the integration of remote sensing (RS) data, the national forest inventory data and regional forest maps to characterize structural features of the main forest species. Gross primary production (GPP) is simulated over the period 2005-2019 and the multiyear predictive capability of the model in simulating GPP is evaluated both aggregated as at species-level by means of independent multiple data sources based on different RS-based products. We show that the model is able to reproduce most of the spatial (∼2800 km2) and temporal (32 years in total) patterns of the observed GPP at both seasonal, annual and interannual time scales, even at the species-level. These new very promising results open the possibility of applying the 3D-CMCC- FEM confidently and robustly to investigate the forests’ behavior under climate and environmental variability over large areas across the highly variable ecological and bio- geographical heterogeneity of the Mediterranean region.
... Tropospheric ozone (O 3 ) is recognized as one of the most harmful air pollutants for plant growth and development (Grulke and Heath, 2020). Since global-scale background O 3 concentrations have been increasing, O 3 pollution is a significant issue for terrestrial ecosystems now and in the near future De Marco et al., 2022). Although many experimental studies have been conducted in controlled chambers and free-air O 3 facilities, they mainly focused on boreal and/or temperate plant species (Grulke and Heath, 2020), and to best of our knowledge, only a few studies were conducted about the responses to O 3 for plants of arid or semi-arid climates such as date palm (Phoenix ☆ This paper has been recommended for acceptance by Parvaiz Ahmad. ...
Article
Ozone (O3) pollution is harmful to plants and ecosystems. Several chemicals have been evaluated to protect plants against O3 deleterious effects. However, they are not adequately efficient and/or the environmental safety of their application is questioned. Hence, new chemicals that provide sufficient protection while being safer for environmental application are needed. This study investigates the response of two O3-sensitive plant species (Phaseolus vulgaris L. cv. Pinto and Nicotiana tabacum L. cv. Bel-W3) leaf-sprayed with deionized water (W, control), ethylenediurea (EDU, 1 mM) or melatonin at lower (1 mM) or higher (3 mM) concentrations (Mel_L and Mel_H, respectively), and then exposed to a square wave of 200 ppb O3, lasting 1 day (5 h day−1) for bean and 2 days (8 h day−1) for tobacco. In both species, the photosynthetic activity of O3-exposed plants was about halved. O3-induced membrane damage was also confirmed by increased malondialdehyde (MDA) byproducts compared to control (W). In EDU- and Mel-treated bean plants, the photosynthetic performance was not influenced by O3, leading to reduction of the incidence and severity of O3 visible injury. In bean plants, Mel_L mitigated the detrimental effect of O3 by boosting antioxidant enzyme activities or osmoprotectants (e.g. abscisic acid, proline, and glutathione transferase). In Mel_L-sprayed tobacco plants, O3 negatively influenced the photosynthetic activity. Conversely, Mel_H ameliorated the O3-induced oxidative stress by preserving the photosynthetic performance, preventing membrane damage, and reducing the visible injuries extent. Although EDU performed better, melatonin protected plants against O3 phytotoxicity, suggesting its potential application as a bio-safer and eco-friendlier phytoprotectant against O3. It is worth noting that the content of melatonin in EDU-treated plants remained unchanged, indicating that the protectant mode of action of EDU is not Mel-related.
Article
Full-text available
Significance Tree diversity is fundamental for forest ecosystem stability and services. However, because of limited available data, estimates of tree diversity at large geographic domains still rely heavily on published lists of species descriptions that are geographically uneven in coverage. These limitations have precluded efforts to generate a global perspective. Here, based on a ground-sourced global database, we estimate the number of tree species at biome, continental, and global scales. We estimated a global tree richness (≈73,300) that is ≈14% higher than numbers known today, with most undiscovered species being rare, continentally endemic, and tropical or subtropical. These results highlight the vulnerability of global tree species diversity to anthropogenic changes.
Article
Full-text available
Forests are increasingly subject to a number of disturbances that can adversely influence their health. Remote sensing offers an efficient alternative for assessing and monitoring forest health. A myriad of methods based upon remotely sensed data have been developed, tailored to the different definitions of forest health considered, and covering a broad range of spatial and temporal scales. The purpose of this review paper is to identify and analyse studies that addressed forest health issues applying remote sensing techniques, in addition to studying the methodological wealth present in these papers. For this matter, we applied the PRISMA protocol to seek and select studies of our interest and subsequently analyse the information contained within them. A final set of 107 journal papers published between 2015 and 2020 was selected for evaluation according to our filter criteria and 20 selected variables. Subsequently, we pair-wise exhaustively read the journal articles and extracted and analysed the information on the variables. We found that (1) the number of papers addressing this issue have consistently increased, (2) that most of the studies placed their study area in North America and Europe and (3) that satellite-borne multispectral sensors are the most commonly used technology, especially from Landsat mission. Finally, most of the studies focused on evaluating the impact of a specific stress or disturbance factor, whereas only a small number of studies approached forest health from an early warning perspective.
Article
Full-text available
Climate warming and elevated ozone (eO 3) are important climate change components that can affect plant growth and plant-microbe interactions. However, the resulting impact on soil carbon (C) dynamics, as well as the underlying mechanisms, remains unclear. Here, we show that warming, eO 3 , and their combination induce tradeoffs between roots and their symbiotic arbuscular mycorrhizal fungi (AMF) and stimulate organic C decomposition in a nontilled soybean agroecosystem. While warming and eO 3 reduced root biomass, tissue density, and AMF coloniza-tion, they increased specific root length and promoted decomposition of both native and newly added organic C. Also, they shifted AMF community composition in favor of the genus Paraglomus with high nutrient-absorbing hyphal surface over the genus Glomus prone to protection of soil organic C. Our findings provide deep insights into plant-microbial interactive responses to warming and eO 3 and how these responses may modulate soil organic C dynamics under future climate change scenarios.
Article
Full-text available
1. Understanding the processes that shape forest functioning, structure, and diversity remains challenging, although data on forest systems are being collected at a rapid pace and across scales. Forest models have a long history in bridging data with ecological knowledge and can simulate forest dynamics over spatio-temporal scales unreachable by most empirical investigations. 2. We describe the development that different forest modelling communities have followed to underpin the leverage that simulation models offer for advancing our understanding of forest ecosystems. 3. Using three widely applied but contrasting approaches –­ species distribution models, individual-­based forest models, and dynamic global vegetation models –­ as examples, we show how scientific and technical advances have led models to transgress their initial objectives and limitations. We provide an overview of recent model applications on current important ecological topics and pinpoint ten key questions that could, and should, be tackled with forest models in the next decade. 4. Synthesis. This overview shows that forest models, due to their complementarity and mutual enrichment, represent an invaluable toolkit to address a wide range of fundamental and applied ecological questions, hence fostering a deeper understanding of forest dynamics in the context of global change.
Article
Surface ozone (O3) is a threat to forests by decreasing photosynthesis and, consequently, influencing the strength of land carbon sink. However, due to the lack of continuous surface O3 measurements, -p observational-based assessments of O3 impacts on forests are largely missing at hemispheric to global scales. Currently, some metrics are used for regulatory purposes by governments or national agencies to protect forests against the negative impacts of ozone: in particular, both Europe and United States (US) makes use of two different exposure-based metrics, i.e. AOT40 and W126, respectively. However, because of some limitations in these metrics, a new standard is under consideration by the European Union (EU) to replace the current exposure metric. We analyse here the different air quality standards set or proposed for use in Europe and in the US to protect forests from O3 and to evaluate their spatial and temporal consistency while assessing their effectiveness in protecting northern-hemisphere forests. Then, we compare their results with the information obtained from a complex land surface model (ORCHIDEE). We find that present O3 uptake decreases gross primary production (GPP) in 37.7% of the NH forested area of northern hemisphere with a mean loss of 2.4% year–1. We show how the proposed US (W126) and the currently used European (AOT40) air quality standards substantially overestimate the extension of potential vulnerable regions, predicting that 46% and 61% of the Northern Hemisphere (NH) forested area are at risk of O3 pollution. Conversely, the new proposed European standard (POD1) identifies lower extension of vulnerability regions (39.6%).
Article
Ground‐level ozone (O 3 ) pollution frequently coincides with the deposition of anthropogenic nitrogen (N), and both factors can influence the structure and functionality of both above‐ and belowground ecosystems. Elevated O 3 levels have been shown to adversely impact plants in many prior reports, but the interacting effects of high O 3 levels and N addition on exposed plants remain to be clearly defined, and the changes in rhizosphere microbial community composition in this context have yet to be studied. The direct and indirect mechanisms and interactions among plants, microbes, and the soil that shape these O 3 and N responses are also poorly understood. Herein, we explored the interactive effects of O 3 exposure (five levels) and soil N (four levels) on the composition of rhizosphere soil microbial communities associated with poplar trees ( Populus euramericana cv. ‘74/76’). In these analyses, exposure to higher levels of O 3 was linked to significant decreases in bacteria, fungi, arbuscular mycorrhizal fungi, and to a reduction in the ratio of fungi‐to‐bacteria, whereas soil N addition had no impact on these parameters. No interactive effects between O 3 and N were observed in the context of alterations in soil microbial community composition, and equivalent performance was observed for concentration‐based (AOT40, cumulative exposure to hourly O 3 concentrations >40 ppb) and flux‐based [POD 1 and POD 7 , cumulative stomatal uptake of O 3 > 1 or 7 nmol O 3 m ⁻¹ PLA (projected leaf area) s ⁻¹ ] dose–response analyses. Structural equation modelling revealed that changes in the composition of the microbial community were attributable to changes in soil pH but unrelated to plant characteristics. Overall, these findings indicated that increased O 3 levels can induce soil alkalinisation and thereby influence soil microbial communities such that soil pH is a reliable predictor of O 3 pollution‐related changes in these communities. Highlights Elevated O 3 exhibited an overall negative influence on microorganisms. AOT40‐, POD 1 ‐ and POD 7 ‐based dose–responses performed equally well. Soil pH is a reliable predictor of O 3 pollution‐related changes in microbial communities. N addition failed to affect O 3 effects on soil microbial community.
Article
Anthropogenic activities have dramatically altered the global nitrogen (N) cycle. Atmospheric N deposition, primarily from combustion of biomass and fossil fuels, has caused acidification of precipitation and freshwater and triggered intense research into ecosystem responses to this pollutant. Experimental simulations of N deposition have been the main scientific tool to understand ecosystem responses, revealing dramatic impacts on soil microbes, plants, and higher trophic levels. However, comparison of the experimental treatments applied in the vast majority of studies with observational and modelled N deposition reveals a wide gulf between research and reality. While the majority of experimental treatments exceed 100 kg N ha⁻¹ y⁻¹, global median land surface deposition rates are around 1 kg N ha⁻¹ y⁻¹ and only exceed 10 kg N ha⁻¹ y⁻¹ in certain regions, primarily in industrialized areas of Europe and Asia and particularly in forests. Experimental N deposition treatments are in fact similar to mineral fertilizer application rates in agriculture. Some ecological guilds, such as saprotrophic fungi, are highly sensitive to N and respond differently to low and high N availability. In addition, very high levels of N application cause changes in soil chemistry, such as acidification, meaning that unrealistic experimental treatments are unlikely to reveal true ecosystem responses to N. Hence, despite decades of research, past experiments can tell us little about how the biosphere has responded to anthropogenic N deposition. A new approach is required to improve our understanding of this important phenomenon. First, characterization of N response functions using observed N deposition gradients. Second, application of experimental N addition gradients at realistic levels over long periods to detect cumulative effects. Third, application of non-linear meta-regressions to detect non-linear responses in meta-analyses of experimental studies.
Article
The ongoing climate change have multi-faceted effects not only on metabolism of plants, but also on the soil properties and mycorrhizal fungal community. Under climate change the stability of the entire forest ecosystems and the carbon balance depend to a large degree on the interactions between trees and mycorrhizal fungi. The main drivers of climate change are CO<sub>2</sub> enrichment, temperature rise, altered precipitation patterns, increased N deposition, soil acidification and pollutants, ecosystem fragmentation and habitat loss, and biotic invasion. These drivers can impact mycorrhizal community directly and indirectly. We discussed the influence of each driver on mycorrhizal community and outlined how mycorrhizas play an important role in the resilience and recovery of forest ecosystems under climate change, by mitigating detrimental effects of CO<sub>2</sub> enrichment, temperature rise, drought, lack of nutrients, soil acidification, pollutants, pests, and diseases. Conservation of the overall biodiversity in forest ecosystems as well as providing the most favourable conditions for the development of mycorrhizae can contribute to increasing the resilience of forest ecosystems to climate change.
Article
Biologically meaningful and cost-effective indicators are needed for assessing and monitoring the impacts of tropospheric ozone (O3) on vegetation and are required in Europe by the National Emission Ceilings Directive (2016). However, a clear understanding on the best suited indicators is missing. The MOTTLES (MOnitoring ozone injury for seTTing new critical LEvelS) project set up a new generation network for O3 monitoring in forest plots in order to: 1) estimate the stomatal O3 fluxes (Phytotoxic Ozone Dose above a threshold Y of uptake, PODY); and 2) collect visible foliar O3 injury, both within the forest plot (ITP) and along the Light Exposed Sampling Site (LESS) along the forest edge. Nine forest sites at high O3 risk were selected across Italy over 2017 − 2019 and significant correlations (p < 0.05) were found between the percentage of symptomatic plant species within the LESS, and POD1 (PODY, with Y = 1 nmol O3 m−2 s−1) calculated for mixed forest species (r = 0.53) and with the occurrence and severity of visible foliar O3 injury on the dominant species in the plots (r = 0.65). A generic flux-based critical level for mixed forest species was derived within the LESS and it was recommended using 11 mmol m−2 POD1 as the critical level for forest protection against O3 injury, similar to the critical level obtained in the ITP (12 mmol m−2 POD1). It was concluded that the frequency of symptomatic plant species within a LESS is a suitable and effective plant-response indicator of phytotoxic O3 levels in forest monitoring. LESS is a non-destructive, less complex and less time-consuming approach compared to the ITP for monitoring foliar O3 injury in the long term. Assessing visible foliar O3 injury in the ITP might only underestimate the O3 risk assessment at individual sites. These results are biologically meaningful and useful to monitoring experts and environmental policy makers.
Article
Given the high ozone concentrations observed in the Mediterranean region during summer, it is crucial to extend our knowledge on the potential ozone impacts on forest health with in situ studies, especially to protect typical endemic forests of the Mediterranean basin. This study is focused on ozone measurements and exposures over the Eastern Adriatic coast and on the calculation of different O3 metrics, i.e., accumulated exposure AOT40 (AOT40dir, AOT40ICP, AOT40pheno) and stomatal O3 fluxes with an hourly threshold of uptake (Y) to represent the detoxification capacity of trees (PODY, with Y = 0, 1, 2 nmol O3 m⁻² s⁻¹) used for forest protection. Finally, we provide an assessment of the relationships between the forest response indicators and environmental variables. Passive ozone measurements and monitoring of forest health indicators, namely growth and crown defoliation, were performed for Quercus ilex, Quercus pubescens, Pinus halepensis, and Pinus nigra forests. Results showed that, for all the analysed species, ozone levels were close to reached the upper plausibility limits for passive monitoring of air quality at forest sites (100 ppb), with the highest values found on P. halepensis in the summer period. O3 metrics based on exposure were found to be higher in pine plots than in oak plots, while the highest values of uptake-based metrics were found on P. nigra. Regarding relationships between environmental variables and forest-health response indicators, the crown defoliation was significantly correlated with the soil water content at various depth while the tree growth was correlated with the different O3 metrics. The most important predictors affecting tree growth of Q. pubescens and Q. ilex were AOT40pheno and AOT40dir and POD0 for P. nigra.