4,705
Views
0
CrossRef citations to date
0
Altmetric
Review

Air pollution and climate change impact on forest ecosystems in Asian region – a review

, &
Article: 2090448 | Received 09 Jun 2021, Accepted 09 Jun 2022, Published online: 19 Jul 2022

ABSTRACT

Forests are complex ecosystems comprising various trophic levels responsible for carrying out various biogeochemical processes and providing ecosystem services. However, forests in Asia are doubly challenged by climate change and air pollution. The rapidly changing air quality, with increasing concentration of greenhouse gases (GHGs), trace gases, volatile organic compounds (VOCs), and ozone (O3) also causes global warming leading to climate change, thus jointly creating a challenging condition for the forest ecosystem. The impact on forest ecosystems of the two anthropogenic stressors, viz., climate change and air pollution, requires global attention. These two stressors have been widely studied separately but their combined impact on the forest ecosystem has not been studied extensively, particularly in the Asian region. In this review article, we attempt to explore the importance of interlinking air pollution and climate change impact on Asian forests, by studying the decline of different forest types as a background and markers of forest ecosystem degradation. Our main aim is to understand and summarise the past and ongoing research in this area and to facilitate researchers and policymakers to upgrade their research, policies, and management strategies in the area of integration of air pollution and climate change impact on forest ecosystems in the Asian region.

Introduction

Climate change and air pollution are two intertwined components that have a strong influence on the forest. Matyssek et al.(Citation2012) described air pollution as a component of climate change. In the atmosphere, anthropogenic and biogenic emissions play an important role in causing air pollution and climate change. Due to chemical reactions occurring in the atmosphere, a number of secondary air pollutants are formed. O3 is one of the secondary air pollutants which is highly phytotoxic in nature. Biogenic volatile organic compounds (BVOCs) particularly emitted from trees have a significant contribution to the formation of O3 (Ng et al. Citation2008; Tasoglou and Pandis Citation2015). However, the roles of BVOCs and climate for future O3 formation are still unclear. Climate change may well increase foliage in most regions, particularly in boreal and temperate regions and this alongwith direct temperature effects can be responsible to accelerate the increase in BVOC emissions in the future and therefore further contributes to an increase in O3 production (Bai and Hao Citation2018). Carbon dioxide (CO2) is also released from natural as well as anthropogenic combustion-related sources. Thus, O3 and CO2 both are important greenhouse gases that play a vital role in increasing the earth surface temperature, change in weather, and precipitation patterns that ultimately cause climate change (Karnosky et al. Citation2002; IPCC Citation2007; Shi et al. Citation2016; Guevara-Ochoa et al. Citation2020; Gruda et al. Citation2021). Moreover, nitric acid and sulphuric acid are also formed from oxides of nitrogen (NOx) and sulphur dioxide (SO2) respectively and play an important role in acid deposition. Acid deposition processes hamper the environment particularly, forest ecosystem health. Thus, the changing climate also affects biogeochemical cycles that further worsen the atmosphere and provide an environment conducive to the formation of secondary pollutants (Ning et al. Citation2020; Sonwani and Maurya Citation2018). This has all been explained clearly in the schematic representation of the interlink between air pollution and climate change in .

Figure 1. Schematic representation of interlink between air pollution and climate change.

Figure 1. Schematic representation of interlink between air pollution and climate change.

Climate change and air pollution are serious threats to forests. Forest ecosystems are a function of biological (plants, animals, micro-organisms), physical (soil, water, temperature, light, precipitation, etc.), and chemical (organic and inorganic constituents) factors that interact with one another to form a self-sustaining ecosystem on their own. Globally, forests cover 30% of the terrestrial land, act as a carbon sink and store 80% of all aboveground carbon and more than 70% of all soil organic carbon (Joshi and Singh Citation2020). The combined effects of air pollution and climate change have the capacity to turn a forest from a C sink to a C source that can further change the dynamics of atmospheric conditions. The key determinants for evaluating the integration of air pollution and climate change are air quality with respect to CO2 concentration (Chatterjee et al. Citation2018), nitrogen (N) deposition (Jenkins Citation2021), O3 exposure vis-a-vis precursor availability (VOCs, NOx) (Gong et al. Citation2021); and climate change impacts through altered insulation, air temperature and precipitation (Vandemeulebroucke et al. Citation2021). The relationship between forests and their interaction with air pollution and climate change follows a circular path. Air pollution and climate change together alter the components of forests like C-sequestration, photosynthesis and transpiration rates, biodiversity, species composition, etc., which thereafter alters the environmental processes. The forest in itself is a complete ecosystem, providing fibre, fuel, food, wood and other non-wood products, freshwater, ecosystem services like air quality, climate regulation, water regulation, erosion regulation, pollination, and natural hazard regulation. The two stressors (climate change and air pollution) have the potential to alter forest product management and ecosystem functions (Lu et al. Citation2019). The impact on forests can be both positive, such as an increase in forest growth from the CO2 fertilization effect i.e., increase in CO2 concentration also stimulates the gross primary productivity and leaf growth, and therefore having positive feedback on BVOC emissions by increasing the emitting biomass (Tao and Jain Citation2005; Dawson et al. Citation2009; Feng et al. Citation2014) and negative, such as an increase in insects and pathogens, drought affecting tree health, disturbance in biogeochemical cycles, and altered precipitation patterns (Lloyd and Bunn Citation2007). Asian forests are at high risk due to high pollutant emissions and global warming due to developmental activities in the late 20th century (Zafar et al.Citation2019). Moreover, climate change factors like GHGs including CO2, methane (CH4), nitrous oxide (N2O), and O3 and extreme events (drought, floods, heat stress, extreme precipitation) have also detrimental effects on forest health (Wang et al. Citation2020). For example, in China, high O3 events are mostly related to the subsidence ahead of a forthcoming tropical cyclone and as a consequence, changes recorded in tropical cyclone numbers and trajectories in the future will efficiently affect local O3 air quality and in turn, it affects the forest health (Shu et al.Citation2016; Lam et al. Citation2018). Biomass burning is another factor that aggravates air pollution and climate change. Though it can be both man-made and natural, the latter accounts for only 10%. Equatorial Asia is responsible for 11% of the global biomass burning annually, releasing radiative gases like CO2, carbon monoxide (CO), NOx, CH4, VOC, and particulate matter (Pei et al. Citation2021). Increasing phytotoxicity has been reported in many parts of East and Southeast Asia due to the rising pollution load (Agathokleous, Kitao, and Kinose Citation2018; Saxena and Sonwani Citation2020). Asian forests are mostly seen to be threatened by elevated concentrations of O3 (Sitch et al. Citation2007; Li et al. Citation2018). The northern hemisphere has seen O3 increase by 1–5ppb/decade since the 1950s, with the highest concentration in Asia (Emberson Citation2020). O3 also has an interesting phenomenon that it itself interacts with vegetation (like forests) to pose an impact on total surface fluxes of BVOCs and O3. Increasing stomatal uptake of ozone affects the photosynthesis process, resulting in decrease in leaf area index (LAI), gross primary productivity, and transpiration, eventually constituting negative feedback via BVOC emissions. Moreover, a decrease in LAI slows the dry deposition velocity of ozone via stomatal uptake, constituting positive feedback (Franz et al. Citation2017; Zhou et al. Citation2018). Nitrogen deposition and sulphur (S) also poses a threat to the forest ecosystem (Etzold et al. Citation2020), in gaseous forms, through acid rains (Chuman et al. Citation2021) and heavy metal toxicity (Hahn et al.Citation2019). The Acid Deposition Monitoring Network in East Asia (EANET) has been monitoring acid deposition and air quality in Asian countries with respect to forests since 1998 (Itahashi et al. Citation2020). The 2016 and 2019 EANET reports revealed that, as compared to Europe and the United States, wet depositions of S in large cities of East Asia were much higher (EANET Citation2016, Citation2019). S concentration was also reported higher in Jinyushan (China) and Ulaanbaatar (Mongolia). Wet deposition of non sea-salt (nss)-S in Japan (8.01 kg S/ha/year) was thrice the amount recorded in Europe (2.57 kg S/ha/year) and the United States (2.85 kg S/ ha/year, NADP) (EMEP Citation2004, Citation2020). Reviews in the area of air pollution and forest decline solely in Asia were not reported so far, till the study of Takahashi et al. (Citation2020). This was due to such studies being published in domestic journals and languages. They chose data particularly from EANET because it uses unified protocols and offers access to transparent and verifiable data.

The influences of air pollution and climate change on forest health have always been segregated (Xiang et al. Citation2020; Liu et al. Citation2021; Joseph Citation2021). However, it is important to study the integrated effects because of the synergistic or antagonistic relation between the two stressors (Matyssek et al. Citation2017). Thus, in this article, we have attempted to adopt a holistic approach to understand the integrated effects of climate change and air pollution on the forest ecosystem, identify specific markers of forest ecosystem degradation, forest types and their health status in Asian region and identify the research gaps to provide relevant recommendations for further research in the field.

Air pollution and climate change linkage in forest

Earlier, research on the forest ecosystem was confined to the impact of air pollutants on trees (Smith Citation2012). However, little research has been conducted on the forest ecosystem, compared to tree health-related research. While studying forest health, their integrated interactions with VOC emissions, CO2, O3, atmospheric deposition, S and N deposition, nutrient and water cycle, ecosystem and social aspects must also be investigated (Serengil et al. Citation2011). For example, elevated CO2 concentrations are very much responsible to reduce stomatal conductance in tree/plant species and have been expected to decrease O3 impacts by limiting stomatal uptake of O3. On the other hand, this concept is more complex in the future environment with multiple stress factors (Wagg et al. Citation2013; Izuta Citation2017). Another example is N deposition which has varying effects on N2O emission in different natural ecosystems i.e., N2O emissions have recorded an increase in its concentrations in an alpine meadow, and arctic tundra (Jiang et al. Citation2010; Zhang et al. Citation2017) while reductions in temperate forests (Skiba et al. Citation1999; Geng et al. Citation2019). However, the mechanism for divergent effects of N deposition and N2O emission remain unclear. Moreover, acid deposition is also one of the important threats to the environment and is responsible for high pollution load in the atmosphere, soil and aquatic ecosystem due to the formation of sulphuric acid and nitric acid vapour in the atmosphere (Gao et al. Citation2017). It further increases nitrogen deposition and the production of nitrogen compounds (HNO3, NO, NH3) in soil (Sanderson et al. Citation2006; Gao et al. Citation2017). The integrated effects of climate change and air pollution threaten the forest ecosystem by changing the soil processes, hampering tree growth, altering species composition and distribution, making plants susceptible to stressors and increasing the probability of wildfires (Bytnerowicz, Omasa, and Paoletti Citation2007; Peñuelas and Sardans Citation2021). In view of the integrated impacts of air pollution and climate change on forest ecosystem, natural and anthropogenic sources of greenhouse gases with other air pollutants, atmospheric chemistry of pollutants, the interaction of the different air pollutants with different plant tissues, and related impacts on plant physiological process and the soil chemistry of chemicals in presence of microorganisms are summarized respectively in . CO2, N2O, and CH4 are emitted from some natural sources like grasslands, wetlands, and cow respiration, whereas, combustion activities, fossil fuel emissions, and motor vehicle emissions are some of the anthropogenic sources which release an enormous amount of the GHGs into the atmosphere and participate in climate change. In case of natural sources, not only soils and herbaceous plants but trees are also an important source of N2O and CH4 to the atmosphere (Machacova et al. Citation2019). The trace gas exchange capacity of trees and their role in the ecosystem, where N2O and CH4 exchange observed, nevertheless, vary significantly among tree individuals, tree species, kinds of forest ecosystems and climatic zones, and they are also dependent on other factors also like tree size, age and health conditions (Barba et al. Citation2019). Thus, air pollutants participate in atmospheric transformation reactions and create several adverse impacts related to plant physiology such as leaf damage, stomata closure, chloroplast injury, lower ROS production, leaf-loss by premature and accelerated senescence, and impedes phloem in plants, resulting in deterioration in the forest ecosystem health and forest decline.

Figure 2. Integrated effects of air pollution and climate change on forests.

Figure 2. Integrated effects of air pollution and climate change on forests.

The higher the temperature and precipitation, the faster the weathering rate. According to Gislason et al. (Citation2009) a significant linear positive correlation was found between mean annual temperature and weathering. For each degree of temperature increase, the weathering flux was found to increase about 30%. The elevated temperature, along with VOC and N oxides, leads to an increase in O3 concentration (The Royal Society Citation2008). Since the 1990s, anthropogenic O3 precursor decreased in North America and Europe, but increased in East Asia (NOx: 4.3%year−1, VOCs: 2.3% year−1) (Sicard Citation2021). In light of this, BVOC emissions have increased by 55.38% due to biomass growth and climate change, of which changing pattern of biomass dominates the interannual variations in emission in the past four decades in China and will affect contributions to O3 formation and which in turn affect the forest/vegetation health (Li and Xie Citation2014; Li et al. Citation2021). O3 is regarded as the main phytotoxic agent having an impact on plant anatomy, ultra-structural and photosynthetic apparatus, tree health, and growth (Sicard et al. Citation2011). Some climate change parameters like elevated temperature make trees more sensitive to air pollutants like SO2 and O3, whereas water stress and elevated CO2 parameters trigger stomatal closure, which in turn helps trees to deposit fewer air pollutants (Ngarambe et al. Citation2021). Moreover, it has also been found by Folberth et al. (Citation2006) that tropospheric O3 produced by isoprene (most abundant BVOC) adds to the global O3 budget which enhances the global mean radiative forcing by +0.09 W/m2 and in tropics, where BVOCs are found in large amount, the radiative forcing can increase to +0.17 W/m2.

Over the past decade, the integrated impacts of air pollution and climate change on forest ecosystems have been explored in many studies and a few are discussed here. Bytnerowicz et al. (Citation2007) reviewed the links between air pollution and climate change and their interactive effects on northern hemisphere forests. They suggested that studying the combined effects is more effective than evaluating one factor alone to enhance the quality of monitoring and better integration of local, national, and global environmental policies. This paper sheds light on the integrated impacts on the Northern Hemisphere for the first time. Bytnerowicz et al. (Citation2013) discussed the status quo and future scope of the collaborative effects of air pollution and climate change on forest ecosystems in the United States. Since the last decade, researchers have lamented the scarcity of studies related to the integrated air pollution and climate change impact on forest ecosystems.

Integrated effects of air pollution and climate change on forests in Asian region

The status of forests in the upcoming years can be predicted with the help of climate, vegetation, forest economics models and socio-economic scenarios. The representative concentration pathways (RCP), is a part of the 5th Coupled Model Intercomparison Project (CMIP5) (Taylor et al. Citation2012). The model also used the RCP scenarios of changes in other anthropogenic GHGs such as CH4, N2O, and halocarbons, and anthropogenic aerosols such as sulfate and black carbon. From 2005, the different RCPs are following their own estimates of land use. In the “no-policy” (RCP 8.5) and the “overshoot and decline” (RCP 2.6) pathways, the global managed area continues to increase throughout the 21st century. Thus, RCP 2.6 and RCP 8.5 use climate change models to predict forest decline. According to these, forest area is envisaged to increase approximately by 7% and 5% in 2100 and decline by 3% and 1% in 2200 with respect to the no-climate-change scenario under the RCP 8.5 and RCP 2.6 respectively (Favero et al. Citation2021). The impact of air pollutants and climate change factors on different tree species in Asian region is listed in Supplementary Table 1. In Asia, forests are characterised by tree species richness (Adams and Woodward Citation1989); and are mainly populated by Siberian larch (Larix sibirica), Japanese larch (L. kaempferi), Scots pine (Pinus sylvestris), Dahurian larch (L. gmelinii), Siberian spruce (Picea obovata), and Siberian pine (Pinus sibirica) . However, many hot spots are losing their biodiversity due to high industrialisation, increased vehicular density and over-population that have increased the impact of air pollution and climate change especially in Asia, particularly Southeast Asia (Myers et al. Citation2000; Qu et al. Citation2022). There are various studies reported in this context so far like Cai et al. (Citation2021) reported that overpopulation and anthropogenic activities are major high-pressure indicators, responsible for the decline in forest ecosystems of China. Moreover, Rasal et al. (Citation2021) also reported that overpopulation growth is one of the anthropogenic cause for the decline in forest ecosystem health and affect the livelihood of local forest dwellers in India. Apart from above-mentioned trees, it was also mentioned that in several parts of the Asia, species like Larix sibirica, Pinus sylvestris, Picea obovata, and Pinus sibirica (Northern Asia) were disturbed due to the effect of air pollution which alters the internal chemistry and metabolism of these species (Mikhailova et al. Citation2013). Due to the drought, various plants Shorea robusta in India, Abies densa in Bhutan, and the entire forest region of Thotupolakanda Mountain region of Sri Lanka suffered from early sensescence and dieback. Moreover, direct air pollution impacts were also observed in Pinus massoniana and Pinus armandi in China (Bian and Yu Citation1992; Song et al. Citation2016) and Pinus densiflora, P. Koraiensis, and P. Rigida in Korea (Kim and Lee Citation1993) Southeast Asia experienced the highest deforestation among many major tropical regions and is likely to lose about 75% of its original forests by 2100, along with up to 42% of its biodiversity (Sodhi et al. Citation2004). The forest fires of 1997/1998 created massive ecological damage and human suffering. Extended periods of minimal rainfall were the cause of forest fires in Southeast Asia during the Ice Age (Goldammer Citation2007). The fires damaged many regions of Southeast Asia from Papua New Guinea to Malaysia, but Indonesia was greatly impacted. Forest fires are an annual environmental crisis, but the extremely dry conditions caused by the 2015 El Niño are making forest fires more frequent and dangerous (World Resource Institute Citation2017). Reddy et al. (Citation2019) evaluated the data of daily MODIS active fire locations for 15 years (2003–2017) in South Asia and revealed that a total of 522,348 fire points were active across different forest types and the maximum number of forest fires occurred during January-May. Of the seven South Asian countries, Bangladesh has the highest emerging fire hotspot areas (34.2%) in forests, followed by 32.2% in India and 29.5% in Nepal. A study evaluating forest fire in Uttarakhand (a state in India) for the period 2016–2019 reported the highest forest fires in the years 2016 and 2018. The burn area was estimated as 3,438 to 3,567 sq. km. (Bar et al. Citation2020). Moreover, mega deltas of Asia like Ayeyarwady delta of Myanmar, Chao Phraya delta of Thailand, Mekong and Song Hong Deltas of Vietnam, Zhujiang, Changjiang, and Huanghe deltas of China, Lena Delta of Russia, Ganges-Brahmaputra Deltas of India and Bangladesh and Indus Delta of Pakistan are important diverse ecosystems of the region housing a diverse range of forests (IUCN (The World Conservation Union Citation2003; Sanlaville and Prieur Citation2005; Rahmonov et al. Citation2017) and also experience climate change impacts and sea-level rise, resulting in extreme events (Nicholls Citation2004).

Health degradation of different forest ecosystems in Asia

Forests ecosystem in Asia are of three types, viz., temperate, tropical, and boreal. Temperate forests are mostly concentrated in Eastern Asia, covering parts of Korea, China, Russia, and Japan, Tropical forests in Asia are spread across Indonesia, the Malay peninsula (Malaysia, Thailand, Myanmar), India, Laos, and Cambodia and Boreal forests are spread over Russia and Siberia. The forest health in all these regions is discussed below in detail.

Temperate forests

Temperate forests in Asia are prey to anthropogenic activity and global warming that has led to the loss of forest cover and decline in biological diversity. In Japan, Copper (Cu) mining led to elevated SO2 emissions in the late 19th century that killed deciduous broad-leaved forests. In the shrines and parks of the metropolitan area of Tokyo, dieback of Japanese Cedars (Cryptomeria japonica), Zelkova (Z. serrata), and other species have been reported (Yambe Citation1978; Nashimoto, Citation1992). The dieback was investigated by several researchers, resulting in many hypotheses. Sekiguchi et al. (Citation1986) attributed into acid deposition and oxidants, while other groups of Sase et al. (Citation1998a), Citation(1998b) blamed degradation of epicuticular wax, Takamatsu et al. (Citation2001a) proposed clogging of stomata by particulate matter and Ito et al. (Citation2002) suggested soil compaction as the reason for the decline of Cedar trees in Japan. The integrated effects of climate change were considered a reason for forest area decline in Kanto Plain (Takamatsu et al. Citation2001b). It is also seen that when the original dominating species in a forest decline, others start dominating, thus changing the dynamics of tree composition. Temperate forests in Korea have shown a similar decline in forest areas in Seoul and the vicinity of industrial complexes (Choi and Toda Citation2008). Seoul in the 1990s reported a decline in Japanese Red Pine (P. densiflora), Korean Pine (P. Koraiensis), Ginkgo (Ginkgo biloba) (Kim and Lee Citation1993; Lee et al. Citation2019). Kim and Lee (Citation1993) observed that fluorine and S depositions in Black Pine (P. thunbergii) decreased the chlorophyll pigment in the needles, resulting in their decline. In China, the increase in SO2 emissions led to the decline of Masson Pine (P. massoniana) (Houtian and Yichuan Citation1990; Yamaguchi et al. Citation2017). In a pine forest of Nashan, necrosis of needle tips was reported in 85% of the trees, apart from early abscission, thinning of crown, and branch dieback due to elevated concentrations of atmospheric SO2 (Takahashi et al. Citation2020). Rural areas of China also displayed many symptoms of tree decline since the 1970s (Song et al. Citation2016). The National Environmental Protection Agency (NEPA) suggested that more than 40% of the land in China has fallen prey to the effects of acid deposition and has been reduced by about 1.14 × 106 ha due to acid rain (Takahashi et al. Citation2020). It has also been predicted that the Gobi Desert located in Northern China and Southern Mongolia would change to warm temperate desert scrub, and the area of cool temperate desert scrub would transfer to the Khangai Mountains, displacing forest areas due to shifts to the warmer and drier conditions in Mongolia (Ulziisaikhan Citation1996); we currently witness this shift (Rosen et al. Citation2019). Moreover, the world most diverse mountain forests are found among the peaks and valleys of the Eastern Himalayas, ironically, this region is vulnerable to climate change, the main reason being its ecological fragility and economic marginality (CEPF Citation2007).

Tropical forest

Asian tropical forests have undergone the greatest forest loss during 2000–2012, with an annual incremental forest loss of 2101 km2 year−1 and if proper mitigation is not undertaken, tropical trees might become extinct (Deb et al. Citation2018). An acceleration of mean global warming (5–9 0C) in the Himalayan Highlands, Tibetan Plateau, and arid regions of South Asia is a key factor indicating the negative impacts of global warming in the tropical forests. For instance, the climate of northern India is described as gradually transforming from a humid to dry condition. Such change in climate over time alters the forest health. For instance, the Brij region of Uttar Pradesh was full of wet tropical forests, with evergreen trees of Indo-Malayan affinities such as Saraca indica, Mesua ferrea, and Anthocephalus indicus, which can be found currently only in Assam, Bengal, Burma, and the west coast of India (Randhawa Citation1945). Pandey (Citation1978) studied a forest located in the Sonbhadra, Uttar Pradesh, and reported chlorosis and necrosis in tree leaves situated in the vicinity of the Obra thermal power plant. The study revealed an increase in total S and exchangeable potassium and a decrease in total N and available phosphorus close to the source. A decreasing gradient in species richness was also observed while moving towards the source. Dubey et al. (Citation1982) observed a reduction of stem circumference and leaf weight in the Betul forest area situated near Satpura thermal power station. The response of tree species to air pollution varied and Bridelia retusa was the most sensitive species noticed by Trivedi and Dubey (Citation1983). Vij et al. (Citation1983) suggested that air pollution caused loss of photosynthetic pigment in the leaves of Bridelia retusa, Mangifera indica, Tectona grandis, Cassia fistula, and Adina cordifolia. Towards the northeast of the refinery, leaves of Mango and Teak showed chlorotic and necrotic symptoms (Agrawal and Agrawal Citation2000). Karmakar et al. (Citation2019) studied the effects of poor air quality on two tropical forests of West Bengal, India. A clear representation of reduction in carbon sequestration was observed in those severely polluted sites. Singh et al. (Citation2021) investigated the effects of O3 in Leucaena leucocephala, using Free Air Ozone Concentration Enrichment (O3-FACE) facility at intervals of 6, 12, 18, and 24 months. Results showed that net photosynthesis, photosynthetic pigments, lipid peroxidation, stomatal conductance, and transpiration significantly decreased after exposure to O3. Joseph (Citation2021) observed that elevated levels of S and N increase the acidity of the soil, inhibiting tree growth. Most of the studies, focuses on the effect of poor air quality on particular tree species/tropical forests (Karmakar et al. Citation2019; Sahu et al. Citation2020; Singh et al. Citation2021). The impact of air pollution on particular plant/tree species, specifically crops, is extensively studied but research on the overall impact of air pollution on forest ecosystems is still lacking.

Boreal forests

In boreal forests ecosystem, the permafrost dynamics and their interaction with climate is one of the major aspects that need to be taken into consideration. Due to changing environmental conditions such as increasing atmospheric temperature resulting degradation of the continuous permafrost layer to discontinuous permafrost, or, even to swamps and wetland areas (Gauthier et al. Citation2015). For instance, a Larch-dominant taiga community in Eurasia has significant relations with increasing ambient temperatures. Larch taiga plays a vital role in global and regional water–energy–carbon (WEC) cycles and in the climate system (Tanaka et al. Citation2008; Ohta et al. Citation2008). It was also reported that the Larch has adapted better than other species to the permafrost environment and also plays a crucial role in the strength of the permafrost carbon climate feedback (Osawa et al. Citation2009). Moreover, the tolerance of Larch species is also enhanced due to its close association with ectomycorrhiza. Symbiotic ectomycorrhiza improves nutrient uptake and buffers against abiotic and biotic stresses (Choi and Toda Citation2008; Ryu et al. Citation2009). On the other side, the pre-tundra Larch (Larix sp.) forests of Siberia located in the vicinity of Norilsk Industrial Centre (largest city built on permafrost north of the Arctic circle) of Central Siberia were seen to be affected by air pollution, especially in the areas adjacent to the industrial center. Extreme pollution has caused the mass mortality of the Larches since the 1960s, reaching its peak in the 1980s. Norilsk lost about 130 × 103 ha forest cover from 1988 to 1993 (Kiseleva Citation1996). It was also documented that only 23% of Siberian larch trees were survived during 90 years in a remote area near the Norilsk industrial area (Kirdyanov et al. Citation2008). Apart from the Larch, Lichen (Cetraria sp.), Siberian spruce (Picea obovata), and Marsh Labrador Tea (Ledum palustre) were also been affected by this catastrophic damage (Vlasova et al. Citation1992). The Baikal Natural Territory (BNT) situated in Eastern Siberia experienced a decline in forest area due to the 10 major industrial centres located in its vicinity (Takahashi et al. Citation2020). Mikhailova et al. (Citation2013) reported higher concentrations of S and heavy metals like arsenic (As), mercury (Hg), and lead (Pb) in Scots pine needles near the industrial centers, compared to those further away. Air pollution also had physiological impacts in Scots pines like imbalance of nutrients and pigments and degradation of the physio-chemical characteristics of the soil (Mikhailova et al. Citation2020). The industrial center of Bratsk, another region in East Siberia, has an extreme continental climate with widely distributed permafrost and has air pollution similar to the BNT area (Kalugina et al. Citation2015). In 2016, a strong pine decline was reported within a distance of 20–30 km (from the industrial center), the moderate decline within 30–70 km, and a weak decline until 80 km (Takahashi et al. Citation2020). The extensive monitoring in the Siberian region revealed that the loss of trees due to air pollution is dependent on the element and primarily occurs within 80–120 km of the source. However, as compared to temperate and tropical forests, the Boreal forests of the Asian region are in a much healthier state, as during the growing season (May–July), the stress has been reversed caused by climate abundant water availability change. Hence, Boreal forests with sufficient water availability can withstand or even increase growth during periods of rising temperatures (Yim et al. Citation2019).

Markers of forest ecosystem degradation

Tree susceptibility

Climate change and air pollution can together increase the susceptibility of forest ecosystems through natural disturbances like insects, disease, drought, or wind. High deposition of nitrogen causes deficiency of macronutrients like potassium (K), phosphorus (P), magnesium (Mg), and calcium (Ca) that consequently makes plants sensitive to factors like drought or parasite attack (UNECE Citation2005). O3 has the highest phytotoxic potential and it is expected that by 2100, half of the world forests will be exposed to phytotoxic O3 levels (Feng et al. Citation2021, Citation2019). Paoletti et al. (Citation2007) reported the vulnerability of certain tree species to freeze-thaw events due to O3 stress. Elevated O3 levels are known to impede phloem loading, which consequently reduces carbohydrate supply to the roots, reducing their biomass (Arab et al. Citation2021). The sensitivity to O3 is different for different tree species, which can alter the species diversity of forests. Wittig et al. (Citation2009) stated that about 40 ppb of O3 concentrations notably reduced 7% of the total biomass of trees, compared to control trees that demonstrated pre-industrial concentrations. De Marco et al. (Citation2020) reported that the O3 risk to Asian forests was quantified at high spatial resolution for the first time and concluded that O3 impacts on deciduous forests of Asia are more serious than on evergreen forests by using ozone indices like AOT40 and W126. A study in China using AOT40 revealed that O3 reduces annual forest tree biomass growth by 11–13%. Li et al. (Citation2018) conducted a nationwide assessment of O3 risk for forests over China as a whole using AOT40 and found a high risk for forests. Qiao et al. (Citation2019) estimated the risks of O3 exposure to forest health using various O3 exposure indices (M7, W126, AOT40, PODY) and found that 90% of the forests in China were under ozone risk. So, normally, there are two categories of metrics that have been developed, i.e., exposure-based metrics (W126, AOT40 etc.) which focus on the negative effects of ozone on vegetation/forests to be dependent on the canopy O3 concentration only, while the second category is of flux-based metrics, the risk depends on the quantity of O3 entering the leaves through the stomata (or uptake) (Simpson et al. Citation2007). However, there are certain uncertainties in these metrics. For instance, exposure-based metrics do not consider any environmental stress to vegetation, and the risk posed to forests is based on the exposure only without considering any ecophysiological constraint (Anav et al. Citation2016). Hence, to overcome this limitation, flux-based metrics have been developed, which is based on the cumulated stomatal flux higher than a given phytotoxic limit below which vegetation is able to detoxify O3 (Emberson et al. Citation2000). Thus, in some of the studies, PODY is considered better than AOT40/W126 or exposure-based metrics for assessing the impact of O3 on vegetation particularly forests (Anav et al. Citation2022; Mills et al. Citation2018; Simpson et al. Citation2007).

NOx, NH3 and HNO3 vapour are phytotoxic at high concentrations and can impact forests adversely (Bytnerowicz et al. Citation2019). Excessive N deposition reportedly decreases the shoot-to-root length, rendering trees susceptible to windthrow (Liu et al. Citation2021). Broken trees showed wider annual growth rings, suggesting decreased mechanical resistance caused by the combined effect of N deposition, temperature, and CO2 (Matyssek et al. Citation2012). Among all Asian countries, Japan showed a relatively low nitrogen deposition over 2000–2018, except for a rural site, Ijira (Takahashi et al. Citation2020). As reported by Takahashi et al. (Citation2020), the nitrogen deposition in Korea over the period 2000–2018 mostly ranged from 50 mmol m−2y−1 to 100 mmol m−2y−1 and Cambodia and Mongolia showed elevated nitrogen depositions with large variations. Paramee et al. (Citation2005) monitored wet N depositions in Thailand and found 40–41 mmol m−2y−1 in two rural sites, 75–95 mmol m−2y−1 in two urban sites, and 119 mmol m−2y−1 in the capital, Bangkok.

Wildfires are both a cause and result of climate change and are responsible for air quality degradation (Lorenz et al. Citation2010). In 2020, several forest fires occurred in the US, Asia, Europe, Australia, and South America (Kelly et al. Citation2020; Zong, Tian, and Yin Citation2020; Haque et al. Citation2021). According to the National Inter-agency Fire Center, 58,950 wildfires were globally reported in 2020, burning an area of 10.1 million acres, including both forests and croplands (https://www.iii.org/fact-statistic/facts-statistics-wildfires). Reddington et al. (Citation2021) used a combination of regional and global air quality models to examine the impact of forest and vegetation fires on air quality degradation in many parts of Southeast Asia.

Primary productivity

The combined effects of climate change and air pollution, like lengthening of warm periods, elevated CO2 together with the rise in N deposition, have enhanced the net primary productivity (NPP) of forests (Wu et al. Citation2021a, Citation2021b). Rehfuess et al. (Citation1999) posited that an increase in CO2 levels due to climate change, together with increased nitrogen deposition, can increase the NPP of forests by 15–20%. The fertilization effect of CO2 and N and elevated temperatures leads to an increase in the total pool of C and therefore, forests have and will become more effective carbon sinks (Yoshitake and Masuda, Citation1986). Further, the increase in NPP resulted largely from increased leaf biomass and canopy photosynthesis driven by increased CO2 and N deposition (Cannell et al. Citation1998; Ren et al. Citation2011; Zhang et al. Citation2012). This can be well supported by Thomley and Cannell (Citation1996) research work where they showed with the help of modelling studies that in a forest ecosystem how NPP increases due to increased CO2 and is amplified by N supply either from the atmosphere in the form of N deposition or by increased nitrogen fixation. On the other side, Shimizu et al. (Citation2019) utilised a remote sensing-based method combined with Accumulated O3 exposure over a Threshold of 40 ppb (AOT40) and MODIS NPP products to evaluate the impact of regional O3 damage on the NPP. The results showed a 4–48% reduction in NPP for the Kanto area of Japan. The NPP of Korean forests has been reduced by 8.25% due to elevated O3 concentration, and the reduction is predicted to range between 8.47% and 10.55% in the 2050s, and between 5.85% and 11.15% in the 2090s (Park et al. Citation2018). Hence, these studies show that combined effects of climate change and air pollution can increase NPP, however, some air pollutants like O3 can show a reduction in NPP.

Carbon budget and sequestration

Forests are capable of sequestering C and storing it for long periods within their ecosystem, thus regulating atmospheric CO2 concentration. The co-influence of climate change and air pollution can potentially decrease the C sequestration capacity of the forests, making them act as a C source instead of a C sink. Prolonged O3 exposure may suppress the C sequestration capacity, as demonstrated by Karnosky et al. (Citation2003) in their free-air CO2+ O3 experiment. Felzer et al. (Citation2004) used empirical equations derived for trees and incorporated them into a terrestrial Ecosystem Model to explore the effects of O3 on C sequestration and observed that it has been reduced by 18–38 Tg C y−1 since the 1950s. NPP and Net Ecosystem Productivity (NEP) are two factors investigated to understand the C-flux of a forest. Piao et al. (Citation2011) studied NEP in East Asia and found that climate change and rising CO2 have favoured East Asian forests, making them a better C sink. The same study recorded a large difference in NEP estimation in the central parts of Northeast China and Mongolia that experienced warming and drying due to climate change; thus, forests act as strong C sources instead of C sinks. Global warming and increasing N deposition interfere with the carbon (C) cycle. Liu et al. (Citation2021) conducted a field experiment in the boreal forests of the Greater Khingan Mountains of Northeast China by adding N to the forest ecosystem and recording its C sequestration capacity at different N concentrations and found a negative correlation between N concentration and the C sequestration capacity of forests. Xiang et al. (Citation2020) observed that the evergreen broad-leaved forests of China showed low C-sequestration rates owing to past disturbances of tree-species composition, management practices, and site conditions and suggested better forest management practices in Asian forests.

Species diversity

Air pollution stressors, particularly S and N deposition, together with climate change, are seen to significantly control the species composition of a forest (De Vries et al. Citation2003). The change in species composition can also be a result of “species shifting.” The natural ability of trees to adapt may cause them to shift from their original site to a place having favourable growing conditions. Thus, species shifting is a major driver of change in species composition in a forest. For example, globally, about 33.75% of the Larix kaempferi species resides in China. However, under the climate change scenario, Larix is seen to shift northward in Asia, Europe, and China and could adapt or move to higher latitudes/altitudes to mitigate the climate change (Wu et al. Citation2021b). In addition to that, Larix hybrid species, F1 (Larix gmelinii var. japonica x Larix kaempferi; hereafter F1) when comes under close association with ectomycorrhiza (ECM) specialists like Suillus sp, survived better under harsh abiotic stress situations (under elevated CO2/O3 conditions) and hence, its species diversity has increased (Wang et al. Citation2015, Citation2016; Qu et al. Citation2022). Apart from this, air pollution stressors are species-specific – when the species diversity is altered, the sensitivity to air pollution also gets altered. This has consequently changed the sensitivity of an ecosystem due to climate change (Serengil et al. Citation2011). Temperature change is also a driver of alteration in species composition. Certain tree species that require low temperature in winters to promote bud bursting in the subsequent season are affected by global warming and can no longer survive in such regions (Bytnerowicz et al. Citation2007).

Soil processes

Temperature and N deposition alter many soil processes imposing consequences on the entire forest ecosystem. The increased mineralisation increases N availability and leaching, leading to soil acidification. Increased emissions of pollutants like NOx and hydrocarbons, and climate change, worsen the problem of acidification by increasing the production and deposition of HNO3 from NO to soils and by converting NH3 into ammonium sulphate (Sanderson et al. Citation2006; Lee et al. Citation2007). This change is due to more aqueous phase oxidation of sulphur dioxide by hydrogen peroxide, which in turn produces more ammonium sulphate. High N deposition is a threat to the forest ecosystem and is particularly identified by high ammonium/nitrate ratios. Excessive ammonium accumulation in the soil acts as a cation exchanger, expelling other nutritional cations from the soil. Soil microorganisms utilize this excess ammonium, which leads to nitrification and consequent high proliferation of nitrate in the soil, and therefore, respiratory C is lost from the soil, transforming forest ecosystems into C sources rather than sinks. This imbalance of nutrient elements created in the forest ecosystem by N dominance causes tree mortality and suppresses NPP (Paudel et al. Citation2018).

Ectomycorrhiza (ECM) associations in the rhizosphere also play an important role in forest ecosystems. Many of the tree species develop symbiotic relationships with ECM fungi to live in harsh conditions through an efficient nutrient uptake by ECM fungi (Nara et al. Citation2003; Qu et al. Citation2010). But, if a particular tree species has been exposed to high O3 levels, it may weaken the specific rate of inorganic N uptake by roots and reduces the sporocarp production in ECM fungi (Andrew and Lilleskov, Citation2009). However, interestingly, the uptake of essential elements was stimulated by an ECM species-specificity of larch seedlings when exposed to a combination of CO2 and O3 treatment (Wang et al. Citation2015).

Conclusion and future recommendation

Forest ecosystems are fragile and exposed to the changing climate and elevated emissions, thus there is a need for a strategic and integrated approach to identifying the major threats, causes, and factors responsible for forest ecosystem degradation. Asian forests are continuously threatened by climate change and air pollution, driven by rapid economic growth and urbanization. Forest decline was mostly reported between 27°–56°N latitude and 103–145°E longitude in subtropical, temperate, and cool temperate climates, with the highest decline reported in Western Japan and Korea. In Asia, the decline can be mostly attributed to locally situated emission sources like traffic, thermal power plant, electric power plant, other industries, and commercial areas, specifically in the early years of industrial establishment.

In Asia, researchers are still investigating the effect of one or more air pollutants (like O3 , BVOCs, elevated CO2) on tree health, tree susceptibility, shoot to root length, C-sequestration rates, photosynthetic rates, transpiration rates, stomatal conductance, biodiversity loss, and ROS production. For instance, several Asian studies used O3 indices to determine the O3 risk to forests. Even though these studies are advanced and relevant, they focus only on the O3 risk to forests, but are not sufficient to evaluate the impact of air pollution and climate change on forest ecosystem. Hence, we need to acknowledge the importance of forests and invest our best efforts in understanding the interactions between different elements of the forest and the combined action of climate change and air pollution. Our plan of action should focus on integrated monitoring and modeling, deepening our knowledge in the areas covering atmospheric deposition in forests, understanding the physiology, microbiology, and biogeochemical cycle of the forests and the role played by soil and economic aspects in forests. This understanding will need the collaboration of scholars from different fields and will provide us with the needful insights. Thus, the scientific community has to amalgamate every possible stressor to get to the root of the cause and the policymakers must also rethink the existing environmental policies, to bring back healthy forest ecosystems.

Acknowledgments

The authors express their sincere thanks to the editor and reviewers for providing valuable comments and suggestions.

Disclosure statement

No potential conflict of interest was reported by the author(s).

References

  • Adams, J. M., and F. I. Woodward. 1989. “Patterns in Tree Species Richness as a Test of the Glacial Extinction Hypothesis.” Nature 339 (6227, Jun): 699–15. doi:10.1038/339699a0.
  • Agathokleous, E., M. Kitao, and Y. Kinose. 2018. “A Review Study on Ozone Phytotoxicity Metrics for Setting Critical Levels in Asia.” Asian Journal of Atmospheric Environment 12 (1): 1–6. doi:10.5572/ajae.2018.12.1.001.
  • Agrawal, M., and S. B. Agrawal. 2000. “Research on Air Pollution Impacts on Indian Forests.” Air Pollution and the Forests of Developing and Rapidly Industrializing Regions. Report No. 4 of the IUFRO Task Force on Environmental Change, 4:165–187.
  • Anav, A., A. De Marco, C. Proietti, A. Alessandri, A. Dell’Aquila, I. Cionni, P. Friedlingstein, et al. 2016. “Comparing Concentration‐based (AOT40) and Stomatal Uptake (PODY) Metrics for Ozone Risk Assessment to European Forests.” Global Change Biology 22 (4): 1608–1627. doi:10.1111/gcb.13138.
  • Anav, A., A. De Marco, A. Collalti, L. Emberson, Z. Feng, D. Lombardozzi, P. Sicard, et al. 2022. “Legislative and Functional Aspects of Different Metrics Used for Ozone Risk Assessment to Forests.” Environmental Pollution 295: 118690. doi:10.1016/j.envpol.2021.118690.
  • Andrew, C., and E. A. Lilleskov. 2009. “Productivity and Community Structure of Ectomycorrhizal Fungal Sporocarps Under Increased Atmospheric CO2 and O3.” Ecology Letters 12 (8): 813–822. doi:10.1111/j.1461-0248.2009.01334.x.
  • Arab, L., Y. Hoshika, H. Müller, L. Cotrozzi, C. Nali, M. Tonelli, P. Ache, et al. 2021. “Chronic Ozone Exposure Preferentially Modifies Root Rather than Foliar Metabolism of Date Palm (Phoenix Dactylifera) Saplings.” Science of the Total Environment 24 (Sep): 150563.
  • Bai, J. H., and N. Hao. 2018. “The Relationships between Biogenic Volatile Organic Compound (BVOC) Emissions and Atmospheric Formaldehyde in a Subtropical Pinus Plantation in China, Ecol.” Environ. Sci 27: 991–999.
  • Bar, S., B. R. Parida, and A. C. Pandey. 2020. Apr 1. “Landsat-8 and Sentinel-2 Based Forest Fire Burn Area Mapping Using Machine Learning Algorithms on GEE Cloud Platform over Uttarakhand, Western Himalaya.” Remote Sensing Applications: Society and Environment 18: 100324. doi:10.1016/j.rsase.2020.100324.
  • Barba, J., M. A. Bradford, P. E. Brewer, D. Bruhn, K. Covey, J. van Haren, J. P. Megonigal, et al. 2019. “Methane Emissions from Tree Stems: A New Frontier in the Global Carbon Cycle.” New Phytologist 222 (1): 18–28. DOI:10.1111/nph.15582.
  • Bian, Y. M., and S. W. Yu. “Forest Decline in Nanshan, China.” Forest Ecology and Management 51(1–3): 53–59. 1992 Aug 15. doi:10.1016/0378-1127(92)90471-K.
  • Bytnerowicz, A., K. Omasa, and E. Paoletti. “Integrated Effects of Air Pollution and Climate Change on Forests: A Northern Hemisphere Perspective.” Environmental Pollution 147(3): 438–445. 2007 Jun 1. doi:10.1016/j.envpol.2006.08.028.
  • Bytnerowicz, A., M. Fenn, S. McNulty, F. Yuan, A. Pourmokhtarian, C. Driscoll, and T. Meixner. Interactive Effects of Air Pollution and Climate Change on Forest Ecosystems in the United States: Current Understanding and Future Scenarios. Developments in Environmental Science 2013 Jan 1 Vol. 13;333–369 Elsevier
  • Bytnerowicz, A., M. E. Fenn, R. Cisneros, D. Schweizer, J. Burley, and S. L. Schilling. 2019 Mar 1. “Nitrogenous Air Pollutants and Ozone Exposure in the Central Sierra Nevada and White Mountains of California–Distribution and Evaluation of Ecological Risks.” Science of the Total Environment 654: 604–615. doi:10.1016/j.scitotenv.2018.11.011.
  • Cai, X., B. Zhang, and J. Lyu. 2021. “Endogenous Transmission Mechanism and Spatial Effect of Forest Ecological Security in China.” Forests 12 (4): 508. doi:10.3390/f12040508.
  • Cannell, M. G. R., J. H. M. Thornley, D. C. Mobbs, and A. D. Friend. 1998. “UK Conifer Forests May Be Growing Faster in Response to Increased N Deposition, Atmospheric CO2 and Temperature.” Forestry: An International Journal of Forest Research 71 (4): 277–296. doi:10.1093/forestry/71.4.277.
  • CEPF. 2007. Ecosystem Profile: Indo-Burman Hotspot, IndoChina Region. Arlington: WWF, US-Asian Program/CEPF.
  • Chatterjee, A., A. Roy, S. Chakraborty, A. K. Karipot, C. Sarkar, S. Singh, S. K. Ghosh, A. Mitra, and S. Raha. 2018. “Biosphere Atmosphere Exchange of CO2, H2O Vapour and Energy during Spring over a High Altitude Himalayan Forest in Eastern India.” Aerosol and Air Quality Research 18 (10, Oct): 2704–2719. doi:10.4209/aaqr.2017.12.0605.
  • Choi, D. S., and H. Toda. 2008. “Air Pollution and Its Effects on the Pine Species in Korea.” Jpn. J. Atmos. Environ 43: 307–314.
  • Chuman, T., F. Oulehle, K. Zajícová, and J. Hruška. 2021. “The Legacy of Acidic Deposition Controls Soil Organic Carbon Pools in Temperate Forests across the Czech Republic.” European Journal of Soil Science 72 (4, Jul): 1780–1801. doi:10.1111/ejss.13073.
  • Dawson, J. P., P. N. Racherla, B. H. Lynn, P. J. Adams, and S. N. Pandis. 2009. “Impacts of Climate Change on Regional and Urban Air Quality in the Eastern United States: Role of Meteorology.” J. Geophys. Res 114 (D5): D05308. doi:10.1029/2008JD009849.
  • De Marco, A., A. Anav, P. Sicard, Z. Feng, and E. Paoletti. “High Spatial Resolution Ozone risk-assessment for Asian Forests.” Environmental Research Letters 15(10): 104095. 2020 Oct 9. doi:10.1088/1748-9326/abb501.
  • De Vries, W., G. J. Reinds, M. Posch, M. J. Sanz, G. H. Krause, V. Calatayud, and J. P. Renaud, JL. Dupouey, H. Sterba, EM. Vel, M. Dobbertin, P. Gundersen, JCH. Voogd . 2003. “JCH.” Intensive Monitoring of Forest Ecosystems in Europe. Technical Report 2003 United Nations, Economic Commission for Europe and European Commission, Brussels, Geneva, Vol. 1, p 1-161.
  • Deb, J. C., S. Phinn, N. Butt, and C. A. McAlpine. “Climate Change Impacts on Tropical Forests: Identifying Risks for Tropical Asia.” Journal of Tropical Forest Science 30(2): 182–194. 2018 Jan 1. doi:10.26525/jtfs2018.30.2.182194.
  • Dubey, P. S., L. Trivedi, and S. K. Shringi. “Pollution Studies of Betul Forest Area Due to Satpura Thermal Power Station Aerial Discharge Final Report.” DOE project 1982. 19/27:78.
  • EANET 2016. Part I Regional Assessment, Third Periodic Report on the State of Acid Deposition in East Asia. Acid Deposition Monitoring Network in East Asia (EANET), Secretariat for the EANET, United Nations Environment Programme Regional Office for Asia Pacific (UNEP ROAP), Bangkok, Thailand, Network Center for the EANET, Asia Center for Air Pollution Research, Niigata (ACAP), Japan.
  • EANET. 2019. Fourth Report for Policy Makers: ―towards Clean Air for Sustainable Future in East Asia through Collaborative Activities, Secretariat for the Acid Deposition Monitoring Network in East Asia (EANET). Bangkok, Thailand: United Nations Environment Programme Asia and the Pacific Office, United Nations Building.
  • Emberson, L. “Effects of Ozone on Agriculture, Forests and Grasslands.” Philosophical Transactions of the Royal Society A 378(2183): 20190327. 2020 Oct 30. doi:10.1098/rsta.2019.0327.
  • Emberson, L. D., M. R. Ashmore, H. M. Cambridge, D. Simpson, and J. P. Tuovinen. 2000. “Modelling Stomatal Ozone Flux across Europe.” Environmental Pollution 109 (3): 403–413. doi:10.1016/S0269-7491(00)00043-9.
  • EMEP, 2004. EMEP Assessment Part I – European Perspective. Norwegian Meteorological Institute, Oslo, Norway. http://www.emep.int [13 Apr 2010].
  • EMEP, 2020.Assessment of Transboundary Pollution by Toxic Substances: Heavy Metals and POPs, EMEP Status Report 2/2020, European Monitoring and Evaluation Programme (http://en.msceast.org/reports/2_2020.pdf) accessed 14 September 2020.
  • Etzold, S., M. Ferretti, G. J. Reinds, S. Solberg, A. Gessler, P. Waldner, M. Schaub, et al. 2020 Feb 15. “Nitrogen Deposition Is the Most Important Environmental Driver of Growth of Pure, even-aged and Managed European Forests.” Forest Ecology and Management 458: 117762. doi:10.1016/j.foreco.2019.117762.
  • Geng, F, Li, K, Liu, X, Gong, Y, Yue, P, Li, Y. and Han, W, 2019. Long-term effects of N deposition on N2O emission in an alpine grassland of Central Asia. Catena, 182, p.104100.
  • Favero, A., R. Mendelsohn, B. Sohngen, and B. Stocker. “Assessing the long-term Interactions of Climate Change and Timber Markets on Forest Land and Carbon Storage.” Environmental Research Letters 16(1): 014051. 2021 Jan 13. doi:10.1088/1748-9326/abd589.
  • Felzer, B., D. Kicklighter, J. Melillo, C. Wang, Q. Zhuang, and R. Prinn. “Effects of Ozone on Net Primary Production and Carbon Sequestration in the Conterminous United States Using a Biogeochemistry Model.” Tellus B: Chemical and Physical Meteorology 56(3): 230–248. 2004 Jan 1. doi:10.3402/tellusb.v56i3.16415.
  • Feng, Z., A. De Marco, A. Anav, M. Gualtieri, P. Sicard, H. Tian, F. Fornasier, F. Tao, A. Guo, and E. Paoletti. 2019 Oct 1. “Economic Losses Due to Ozone Impacts on Human Health, Forest Productivity and Crop Yield across China.” Environment International 131: 104966. doi:10.1016/j.envint.2019.104966.
  • Feng, Z., E. Agathokleous, X. Yue, E. Oksanen, E. Paoletti, H. Sase, A. Gandin, et al. “Emerging Challenges of Ozone Impacts on Asian Plants: Actions are Needed to Protect Ecosystem Health.” Ecosystem Health and Sustainability 7(1): 1911602. 2021 Jan 1. doi:10.1080/20964129.2021.1911602.
  • Feng, Z., J. Sun, W. Wan, E. Hu, and V. Calatayud. 2014 Oct 1. “Evidence of Widespread ozone-induced Visible Injury on Plants in Beijing, China.” Environmental Pollution 193: 296–301. doi:10.1016/j.envpol.2014.06.004.
  • Folberth, G. A., D. A. Hauglustaine, J. Lathièrel, and F. Brocheton. 2006. “Interactive Chemistry in the Laboratoire de Météorologie Dynamique General Circulation Model: Model Description and Impact Analysis of Biogenic Hydrocarbons on Tropospheric Chemistry.” Atmos. Chem. Phys 6 (8): 2273. doi:10.5194/acp-6-2273-2006.
  • Franz, M., D. Simpson, A. Arneth, and S. Zaehle. 2017. “Development and Evaluation of an Ozone Deposition Scheme for Coupling to a Terrestrial Biosphere Model.” Biogeosciences 14 (1): 45–71. doi:10.5194/bg-14-45-2017.
  • Gao, J., E. Wang, W. Ren, X. Liu, Y. Chen, Y. Shi, and Y. Yang. 2017 Aug 1. Effects of Simulated Climate Change on Soil Microbial Biomass and Enzyme Activities in Young Chinese Fir (Cunninghamia Lanceolata) in Subtropical China. In Acta EcologicaSinica 37(4):272–278.
  • Gauthier, S., P. Bernier, T. Kuuluvainen, A. Z. Shvidenko, and D. G. Schepaschenko. 2015. “Boreal Forest Health and Global Change.” Science 349 (6250): 819–822. doi:10.1126/science.aaa9092.
  • Gislason, S. R., E. H. Oelkers, E. S. Eiriksdottir, M. I. Kardjilov, G. Gisladottir, B. Sigfusson, … N. Oskarsson. 2009. “Direct Evidence of the Feedback between Climate and Weathering.” Earth and Planetary Science Letters 277 (1–2): 213–222. doi:10.1016/j.epsl.2008.10.018.
  • Goldammer, J. G. 2007. “History of Equatorial Vegetation Fires and Fire Research in Southeast Asia before the 1997–98 Episode: A Reconstruction of Creeping Environmental Changes.” Mitigation and Adaptation Strategies for Global Change 12 (1): 13–32. doi:10.1007/s11027-006-9044-7.
  • Gong, C., X. Yue, H. Liao, and Y. Ma. 2021. “A Humidity-Based Exposure Index Representing Ozone Damage Effects on Vegetation.” Environmental Research Letters 16 (4): 044030. doi:10.1088/1748-9326/abecbb.
  • Gruda, N., M. Bisbis, N. Katsoulas, and C. Kittas. “Smart Greenhouse Production Practices to Manage and Mitigate the Impact of Climate Change in Protected Cultivation.” InVIII South-Eastern Europe Symposium on Vegetables and Potatoes 1320, 2021. Sep 24 (pp. 189–196).
  • Guevara-Ochoa, C., A. Medina-Sierra, and L. Vives. 2020 Jun 20. “Spatio-temporal Effect of Climate Change on Water Balance and Interactions between Groundwater and Surface Water in Plains.” Science of the Total Environment 722: 137886. doi:10.1016/j.scitotenv.2020.137886.
  • Hahn, J., B. Mann, U. Bange, and M. Kimmel. 2019 Dec 1. “Horizon-specific Effects of Heavy Metal Mobility on Nitrogen Binding Forms in Forest Soils near a Historic Smelter (Germany).” Geoderma 355: 113895. doi:10.1016/j.geoderma.2019.113895.
  • Haque, M. K., M. A. Azad, M. Y. Hossain, T. Ahmed, M. Uddin, and M. M. Hossain. “Wildfire in Australia during 2019-2020, Its Impact on Health, Biodiversity and Environment with Some Proposals for Risk Management: A Review.” Journal of Environmental Protection 12(6): 391–414. 2021 Jun 11. doi:10.4236/jep.2021.126024.
  • Houtian, L., and L. Yichuan. 1990. “Atmospheric SO2 Pollution and Decline of Masson Pine (Pinus Massoniana) Forest in Nanshan, Chongqing [China].” Acta EcologicaSinica (China), 10(4):305-310.
  • IPCC (2007) Climate Change 2007: Synthesis Report. Summary for Policymakers. WG1, AR4. http://ipcc-wg1.ucar.edu/wg1/wg1-report.html
  • Itahashi, S., B. Ge, K. Sato, J. S. Fu, X. Wang, K. Yamaji, T. Nagashima, et al. “MICS-Asia III: Overview of Model Intercomparison and Evaluation of Acid Deposition over Asia.” Atmospheric Chemistry and Physics 20(5): 2667–2693. 2020 Mar 4. doi:10.5194/acp-20-2667-2020.
  • Ito, E., S. Yoshinaga, Y. Ohnuki, K. Shichi, Y. Matsumoto, and H. Taoda. 2002. “Soil Factors Affecting the Decline of Cryptomeria Japonica Forest in the Kanto Plains, Japan.” Journal of Forest Environment (Japan), 44(2):37-43.
  • IUCN (The World Conservation Union). 2003. “The Lower Indus River: Balancing Development and Maintenance of Wetland Ecosystems and Dependent Livelihoods.” Water and Nature Initiative. 5. Accessed 24.01.07. www.iucn.org/themes/wani/flow/cases/Indus.pdf.
  • Izuta, T. 2017. Air Pollution Impacts on Plants in East Asia, edited by T. Izuta., 322. Tokyo: Springer Japan.
  • Jenkins, H. S. 2021. “Air Pollution and Climate Drive Annual Growth in Ponderosa Pine Trees in Southern California.” Climate 9 (5, May): 82. doi:10.3390/cli9050082.
  • Jiang, C., G. Yu, H. Fang, G. Cao, and Y. Li. 2010. “Short-term Effect of Increasing Nitrogen Deposition on CO2, CH4 and N2O Fluxes in an Alpine Meadow on the Qinghai-Tibetan Plateau, China.” Atmos. Environ 44 (24): 2920–2926. doi:10.1016/j.atmosenv.2010.03.030.
  • Joseph, J. “Environmental and Health Impacts of Air Pollution.” an Evaluation in Indian Context Pollution Research 2021 20 (3): 855–860.
  • Joshi, R., and H. Singh. “Carbon Sequestration Potential of Disturbed and non-disturbed Forest Ecosystem: A Tool for Mitigating Climate Change.” African Journal of Environmental Science and Technology 14(11): 385–393. 2020 Nov 30. doi:10.5897/AJEST2020.2920.
  • Kalugina, O. V., E. N. Taranenko, T. A. Mikhajlova, and O. V. Shergina. 2015. “Technogenic Pollution of Pine Forests by Polycyclic Aromatic Hydrocarbons.” Siberian Journal of Forest Science, 4:1-7.
  • Karmakar, D., T. Ghosh, and P. K. Padhy. 2019 Mar 1. “Effects of Air Pollution on Carbon Sequestration Potential in Two Tropical Forests of West Bengal, India.” Ecological Indicators 98: 377–388. doi:10.1016/j.ecolind.2018.11.014.
  • Karnosky, D. F., K. E. Percy, B. Mankovska, T. Prichard, A. Noormets, R. E. Dickson, E. Jepsen, and J. G. Isebrands. 2003 Jan 1. “Ozone Affects the Fitness of Trembling Aspen”. Developments in Environmental Science 3. 199–209.
  • Karnosky, D. F., K. E. Percy, B. Xiang, B. Callan, A. Noormets, B. Mankovska, A. Hopkin, et al. 2002. “Interacting Elevated CO2 and Tropospheric O3 Predisposes Aspen (Populus Tremuloides Michx.) to Infection by Rust (Melampsora Medusae F. Sp. Tremuloidae).” Global Change Biology 8 (4): 329–338. doi:10.1046/j.1354-1013.2002.00479.x.
  • Kelly, L. T., K. M. Giljohann, A. Duane, N. Aquilué, S. Archibald, E. Batllori, A. F. Bennett, et al. 2020 Nov 20. “Fire and Biodiversity in the Anthropocene”. Science 370. (6519): doi:10.1126/science.abb0355.
  • Kim, M. H., and S. W. Lee. 1993. “Injury Responses of Woody Landscape Plants to Air Pollutants (I)-sulfur and Heavy Metal Content.” Journal of Korean Society of Forest Science 82 (3): 271–282.
  • Kirdyanov, A. V., K. S. Treydte, A. Nikolaev, G. Helle, and G. H. Schleser. 2008. “Climate Signals in Tree‐ring Width, Density and δ13C from Larches in Eastern Siberia (Russia).” Chemical Geology 252 (1–2): 31–41. doi:10.1016/j.chemgeo.2008.01.023.
  • Kiseleva, V., 1996, Environmental Stress to the Siberian Forests: An Overview. IIASA Working Paper (1996), p. 34.
  • Lam, Y. F., H. M. Cheung, and C. C. Ying, 2018, “Impact of Tropical Cyclone Track Change on Regional Air Quality.” The Science of the Total Environment 610: 1347–1355. doi:10.1016/j.scitotenv.2017.08.100.
  • Lee, C. S., S. Jung, B. S. Lim, A. R. Kim, C. H. Lim, and H. Lee. “Forest Decline under Progress in the Urban Forest of Seoul, Central Korea.” InForest Degradation around the World 2019, 1:53-67. May 16. IntechOpen.
  • Lee, C. S., J. S. Moon, and Y. C. Cho. 2007. “Effects of Soil Amelioration and Tree Planting on Restoration of an air-pollution Damaged Forest in South Korea.” Water, Air, and Soil Pollution 179 (1, Feb): 239–254. doi:10.1007/s11270-006-9228-5.
  • Li, P., A. De Marco, Z. Feng, A. Anav, D. Zhou, and E. Paoletti. 2018 Jun 1. “Nationwide ground-level Ozone Measurements in China Suggest Serious Risks to Forests.” Environmental Pollution 237: 803–813. doi:10.1016/j.envpol.2017.11.002.
  • Li, L. Y., and S. D. Xie. 2014. “Historical Variations of Biogenic Volatile Organic Compound Emission Inventories in China, 1981–2003, Atmos.” Environ 95: 185–196.
  • Li, L. Y., B. W. Zhang, J. Cao, S. D. Xie, and Y. Wu. 2021. “Isoprenoid Emissions from Natural Vegetation Increased Rapidly in Eastern China.” Environ. Res 200: 111462. doi:10.1016/j.envres.2021.111462.
  • Liu, G., G. Yan, M. Chang, B. Huang, X. Sun, S. Han, Y. Xing, and Q. Wang. 2021. “Long-term Nitrogen Addition Further Increased Carbon Sequestration in a Boreal Forest.” European Journal of Forest Research 5 (Jun): 1–4.
  • Lloyd, A. H., and A. G. Bunn. “Responses of the Circumpolar Boreal Forest to 20th Century Climate Variability.” Environmental Research Letters 2(4): 045013. 2007 Nov 26. doi:10.1088/1748-9326/2/4/045013.
  • Lorenz, M., N. Clarke, E. Paoletti, A. Bytnerowicz, N. Grulke, N. Lukina, H. Sase, and J. Staelens. 2010. Air Pollution Impacts on Forests in a Changing Climate. International Union of Forest Research Organizations (IUFRO), Vienna, Austria.
  • Lu, S., F. Qin, N. Chen, Z. Yu, Y. Xiao, X. Cheng, and X. Guan. 2019 Sep 1. “Spatiotemporal Differences in Forest Ecological Security Warning Values in Beijing: Using an Integrated Evaluation Index System and System Dynamics Model.” Ecological Indicators 104: 549–558. doi:10.1016/j.ecolind.2019.05.015.
  • Machacova, K., E. Vainio, O. Urban, and M. Pihlatie. 2019. “Seasonal Dynamics of Stem N2O Exchange Follow the Physiological Activity of Boreal Trees.” Nature Communications 10 (1): 4989. doi:10.1038/s41467-019-12976-y.
  • Matyssek, R., G. Wieser, C. Calfapietra, W. De Vries, P. Dizengremel, D. Ernst, Y. Jolivet, et al. 2012 Jan 1. “Forests under Climate Change and Air Pollution: Gaps in Understanding and Future Directions for Research.” Environmental Pollution 160: 57–65. doi:10.1016/j.envpol.2011.07.007.
  • Matyssek, R., A. R. Kozovits, G. Wieser, J. King, and H. Rennenberg. “Woody-plant Ecosystems under Climate Change and Air pollution—response Consistencies across Zonobiomes?” Tree Physiology 37(6): 706–732. 2017 Jun 1. doi:10.1093/treephys/tpx009.
  • Mikhailova, T. A., O. V. Kalugina, and O. V. Shergina. 2020. “Identification the Areas of Concern for Pine Forest and Soil Cover Conditions in the Predbaikalie Region.” InIOP Conference Series: Earth and Environmental Science (Vol. Q25 408, No. 1, p. 012082). IOP Publishing.
  • Mikhailova, T. A., O. V. Kalugina, and O. V. Shergina. 2013. “Phytomonitoring of Air Pollution in the Baikal Region.” Contemporary Problems of Ecology 6 (5, Sep): 549–554. doi:10.1134/S1995425513050119.
  • Mills, G., H. Pleijel, C. S. Malley, B. Sinha, O. R. Cooper, M. G. Schultz, H. S. Neufeld, et al. 2018. “Tropospheric Ozone Assessment Report: Present-day Tropospheric Ozone Distribution and Trends Relevant to Vegetation.” Elem Sci Anth 6 (1). doi:10.1525/elementa.302.
  • Myers, N., R. A. Mittermeier, C. G. Mittermeier, G. A. Da Fonseca, and J. Kent. 2000. “Biodiversity Hotspots for Conservation Priorities.” Nature 403 (6772, Feb): 853–858. doi:10.1038/35002501.
  • Nara, K., H. Nakaya, B. Wu, Z. Zhou, and T. Hogetsu. 2003. “Underground Primary Succession of Ectomycorrhizal Fungi in a Volcanic Desert on Mount Fuji.” New Phytologist 159 (3): 743–756. doi:10.1046/j.1469-8137.2003.00844.x.
  • Nashimoto, M. “Actual State of Plant Decline Phenomena Due to Acid Rain and Others; Decline of Japanese Cedar (Cryptomeria Japonica D. Don) Trees in the Kansai-Setouchi District.” Kogai to Taisaku (Journal of Environmental Pollution Control)Bull.(Japan) 1992. Nov 1 28.14. p. 1335-1339.
  • Ng, N. L., A. J. Kwan, J. D. Surratt, A. W. H. Chan, P. S. Chhabra, A. Sorooshian, H. O. T. Pye, et al. 2008. “Secondary Organic Aerosol (SOA) Formation from Reaction of Isoprene with Nitrate Radicals (NO3).” Atmos. Chem. Phys 8 (14): 4117–4140. doi:10.5194/acp-8-4117-2008.
  • Ngarambe, J., S. J. Joen, C. H. Han, and G. Y. Yun. 2021 Feb 5. “Exploring the Relationship between Particulate Matter, CO, SO2, NO2, O3 and Urban Heat Island in Seoul, Korea.” Journal of Hazardous Materials 403: 123615. doi:10.1016/j.jhazmat.2020.123615.
  • Nicholls, R. J. “Coastal Flooding and Wetland Loss in the 21st Century: Changes under the SRES Climate and socio-economic Scenarios.” Global Environmental Change 14(1): 69–86. 2004 Apr 1. doi:10.1016/j.gloenvcha.2003.10.007.
  • Ning, G., S. H. Yim, Y. Yang, Y. Gu, and G. Dong. 2020 Nov 1. “Modulations of Synoptic and Climatic Changes on Ozone Pollution and Its Health Risks in mountain-basin Areas.” Atmospheric Environment 240: 117808. doi:10.1016/j.atmosenv.2020.117808.
  • Ohta, T., T. C. Maximov, A. J. Dolman, T. Nakai, M. K. van der Molen, A. V. Kononov, … H. Yabuki. 2008. “Interannual Variation of Water Balance and Summer Evapotranspiration in an Eastern Siberian Larch Forest over a 7-year Period (1998–2006).” Agricultural and Forest Meteorology 148 (12): 1941–1953. doi:10.1016/j.agrformet.2008.04.012.
  • Osawa, A., O. A. Zyryanova, Y. Matsuura, and T. Kajimoto. 2009. Permafrost Ecosystems: Siberian Larch Forests, 502. 1st edn ed. Berlin: Springer).
  • Pandey, S. N. Effects of coal smoke and sulphur dioxide pollution on plants. Unpublished thesis, Department of Botany BHU Varanasi. 1978;62.
  • Paoletti, E., A. Bytnerowicz, C. Andersen, A. Augustaitis, M. Ferretti, N. Grulke, … K. Percy. 2007. “Impacts of Air Pollution and Climate Change on Forest ecosystems—emerging Research Needs.” TheScientificWorldJOURNAL 7: 1–8. doi:10.1100/tsw.2007.52.
  • Paramee, S., A. Chidthaisong, S. Towprayoon, P. Asnachinda, V. N. Bashkin, and N. Tangtham. 2005. “Three-year Monitoring Results of Nitrate and Ammonium Wet Deposition in Thailand.” Environmental Monitoring and Assessment 102 (1, Mar): 27–40. doi:10.1007/s10661-005-1593-9.
  • Park, J. H., D. K. Lee, J. Gan, C. Park, S. Kim, S. Sung, T. Y. Jung, and S. C. Hong. 2018. “Effects of Climate Change and Ozone Concentration on the Net Primary Productivity of Forests in South Korea.” Forests 9 (3, Mar): 112. doi:10.3390/f9030112.
  • Paudel, I., M. Halpern, Y. Wagner, E. Raveh, U. Yermiyahu, G. Hoch, and T. Klein. 2018 Apr 1. “Elevated CO2 Compensates for Drought Effects in Lemon Saplings via Stomatal Downregulation, Increased Soil Moisture, and Increased Wood Carbon Storage.” Environmental and Experimental Botany 148: 117–127. doi:10.1016/j.envexpbot.2018.01.004.
  • Pei, F., Y. Zhou, and Y. Xia. 2021. “Assessing the Impacts of Extreme Precipitation Change on Vegetation Activity.” Agriculture 11 (6, Jun): 487 doi:10.3390/agriculture11060487.
  • Peñuelas, J., and J. Sardans. 2021. “Global Change and Forest Disturbances in the Mediterranean Basin: Breakthroughs, Knowledge Gaps, and Recommendations.” Forests 12 (5, May): 603. doi:10.3390/f12050603.
  • Piao, S., P. Ciais, M. Lomas, C. Beer, H. Liu, J. Fang, P. Friedlingstein, et al. “Contribution of Climate Change and Rising CO2 to Terrestrial Carbon Balance in East Asia: A multi-model Analysis.” Global and Planetary Change 75(3–4): 133–142. 2011 Feb 1. doi:10.1016/j.gloplacha.2010.10.014.
  • Qiao, X., P. Wang, J. Zhang, H. Zhang, Y. Tang, J. Hu, and Q. Ying. 2019 Jul 15. “Spatial-temporal Variations and Source Contributions to Forest Ozone Exposure in China.” Science of the Total Environment 674: 189–199. doi:10.1016/j.scitotenv.2019.04.106.
  • Qu, L., K. Makoto, D. S. Choi, A. M. Quoreshi, and T. Koike. 2010. “The Role of Ectomycorrhiza in Boreal Forest Ecosystem.” In A. Osawa, OA. Zyryanova, Y. Matsuura, T. Kajimoto, RW. Wein, Editors. Permafrost Ecosystems, 413–425. Dordrecht: Springer.
  • Qu, L., Y. Wang, O. Masyagina, S. Kitaoka, S. Fujita, K. Kita, A. Prokushkin, and T. Koike, 2022. Larch: A Promising Deciduous Conifer as an eco-environmental Resource. https://www.intechopen.com/online-first/80260.
  • Rahmonov, O., T. Szczypek, T. Niedźwiedź, U. Myga-Piątek, M. Rahmonov, and V. A. Snytko. “The Human Impact on the Transformation of Juniper Forest Landscape in the Western Part of the Pamir-Alay Range (Tajikistan).” Environmental Earth Sciences 76(8): 324. 2017. Apr 1. doi:10.1007/s12665-017-6643-4.
  • Randhawa, M. S. 1945. “Progressive Desiccation of Northern India.” J. Bom. Nat. Hist. Soc 45: 558–565.
  • Rasal, V., M. Dhakad, and D. Khandal. 2021. “Assessing Anthropogenic Pressure of Forest Villages in Semi-arid Forest Ecosystem in Western India Using Cumulative Disturbance Index (CDI).” Arid Ecosyst 11 (4): 351–357. doi:10.1134/S2079096121040107.
  • Reddington, C. L., L. Conibear, S. Robinson, C. Knote, S. R. Arnold, and D. V. Spracklen. 2021. “Air Pollution from Forest and Vegetation Fires in Southeast Asia Disproportionately Impacts the Poor.” GeoHealth 5 (9, Sep): e2021GH000418. doi:10.1029/2021GH000418.
  • Reddy, C. S., N. G. Bird, S. Sreelakshmi, T. M. Manikandan, M. Asra, P. H. Krishna, C. S. Jha, P. V. Rao, and P. G. Diwakar. 2019. “Identification and Characterization of spatio-temporal Hotspots of Forest Fires in South Asia.” Environmental Monitoring and Assessment 191 (3, Dec): 1–7. doi:10.1007/s10661-019-7695-6.
  • Rehfuess, K. E., G. I. Ågren, F. Andersson, M. G. Cannell, A. Friend, I. Hunter, H. P. Kahle, J. Prietzel, and H. Spiecker. 1999. “Relationships between Recent Changes of Growth and Nutrition of Norway Spruce, Scots Pine.” In European Beech Forests in Europe: Recognition. European forest Institute, Finland, 1-94.
  • Ren, W., H. Tian, B. Tao, A. Chappelka, G. Sun, C. Lu, M. Liu, G. Chen, and X. Xu. 2011. “Impacts of Tropospheric Ozone and Climate Change on Net Primary Productivity and Net Carbon Exchange of China’s Forest Ecosystems.” Global Ecology and Biogeography 20 (3): 391–406. doi:10.1111/j.1466-8238.2010.00606.x.
  • Rosen, A. M., T. C. Hart, J. Farquhar, J. S. Schneider, and T. Yadmaa. 2019. “Holocene Vegetation Cycles, land-use, and Human Adaptations to Desertification in the Gobi Desert of Mongolia.” Vegetation History and Archaeobotany 28 (3, May): 295–309. doi:10.1007/s00334-018-0710-y.
  • The Royal Society, 2008. Ground-level Ozone in the 21st Century: Future Trends, Impacts and Policy Implications. Policy Report, October 2008, 15/08.
  • Ryu, K., M. Watanabe, H. Shibata, K. Takagi, M. Nomura, and T. Koike . 2009. “Ecophysiological Responses of the Larch Species in Northern Japan to Environmental Changes as a Basis for Afforestation.” Landscape Ecol Eng 5 (2): 99–106. DOI:10.1007/s11355-009-0063-x.
  • Sahu, C., S. Basti, and S. K. Sahu. 2020. “Air Pollution Tolerance Index (APTI) and Expected Performance Index (EPI) of Trees in Sambalpur Town of India.” SN Applied Sciences 2 (8, Aug): 1–4. doi:10.1007/s42452-020-3120-6.
  • Sanderson, M. G., W. J. Collins, C. E. Johnson, and R. G. Derwent. “Present and Future Acid Deposition to Ecosystems: The Effect of Climate Change.” Atmospheric Environment 40(7): 1275–1283. 2006 Mar 1. doi:10.1016/j.atmosenv.2005.10.031.
  • Sanlaville, P., and A. A. Prieur. 2005. “Middle East, Coastal Ecology and Geomorphology.” In Encyclopedia of Coastal Science, edited by M. L. Schwartz, 71–83. Dordrecht: Springer.
  • Sase, H., T. Takamatsu, T. Yoshida, and K. Inubushi. “Changes in Properties of Epicuticular Wax and the Related Water Loss in Japanese Cedar (Cryptomeria Japonica) Affected by Anthropogenic Environmental Factors.” Canadian Journal of Forest Research 28(4): 546–556. 1998b Apr 1. doi:10.1139/x98-021.
  • Sase, H., T. Takamatsu, and T. Yoshida. 1998a Jan 1. Variation in Amount and Elemental Composition of Epicuticular Wax in Japanese Cedar (Cryptomeria Japonica) Leaves Associated with Natural Environmental Factors. In Canadian Journal of Forest Research 28(1):87–97.
  • Saxena, P., and S. Sonwani. 2020 Oct 1. Remediation of Ozone Pollution by Ornamental Plants in Indoor Environment. In Global Journal of Environmental Science and Management 6(4):497–508.
  • Serengil, Y., A. Augustaitis, A. Bytnerowicz, N. Grulke, A. R. Kozovitz, R. Matyssek, G. Müller-Starck, et al. 2011. “Adaptation of Forest Ecosystems to Air Pollution and Climate Change: A Global Assessment on Research Priorities.” iForest-Biogeosciences and Forestry 4 (2): 44. doi:10.3832/ifor0566-004.
  • Shi, C., M. Kitao, E. Agathokleous, M. Watanabe, H. Tobita, K. I. Yazaki, S. Kitaoka, and T. Koike. 2016. “Foliar Chemical Composition of Two Oak Species Grown in a free-air Enrichment System with Elevated O3 and CO2.” Journal of Agricultural Meteorology 72 (1): 50–58. doi:10.2480/agrmet.D-14-00018.
  • Shimizu, Y., Y. Lu, M. Aono, and K. Omasa. 2019 Nov 15. “A Novel Remote sensing-based Method of Ozone Damage Assessment Effect on Net Primary Productivity of Various Vegetation Types.” Atmospheric Environment 217: 116947. doi:10.1016/j.atmosenv.2019.116947.
  • Shu, L., M. Xie, T. J. Wang, D. Gao, P. L. Chen, and Y. Han, S. Li, B. Zhuang and M. Li . 2016. “Integrated Studies of a Regional Ozone Pollution Synthetically Affected by Subtropical High and Typhoon System in the Yangtze River Delta Region, China.” Atmos Chem Phys 16 (24): 15801–15819. DOI:10.5194/acp-16-15801-2016.
  • Sicard, P., O. Lesne, N. Alexandre, A. Mangin, and R. Collomp. “Air Quality Trends and Potential Health effects–development of an Aggregate Risk Index.” Atmospheric Environment 45(5): 1145–1153. 2011 Feb 1. doi:10.1016/j.atmosenv.2010.12.052.
  • Sicard, P. 2021. “Ground-level Ozone over Time: An observation-based Global Overview.” Current Opinion in Environmental Science & Health 19: 100226. doi:10.1016/j.coesh.2020.100226.
  • Simpson, D., M. R. Ashmore, L. Emberson, and J. P. Tuovinen. 2007. “A Comparison of Two Different Approaches for Mapping Potential Ozone Damage to Vegetation. A Model Study.” Environmental Pollution 146 (3): 715–725. doi:10.1016/j.envpol.2006.04.013.
  • Singh, P., R. Kannaujia, S. Narayan, A. Tewari, P. A. Shirke, and V. Pandey. 2021 Aug 15. “Impact of Chronic Elevated Ozone Exposure on Photosynthetic Traits and anti-oxidative Defense Responses of Leucaena Leucocephala (Lam.) de Wit Tree under Field Conditions.” Science of the Total Environment 782: 146907. doi:10.1016/j.scitotenv.2021.146907.
  • Sitch, S., P. M. Cox, W. J. Collins, and C. Huntingford. 2007. “Indirect Radiative Forcing of Climate Change through Ozone Effects on the land-carbon Sink.” Nature 448 (7155, Aug): 791–794. doi:10.1038/nature06059.
  • Skiba, U., L. J. Sheppard, C. E. R. Pitcairn, S. Van Dijk, and M. J. Rossall. 1999. “The Effect of N Deposition on Nitrous Oxide and Nitric Oxide Emissions from Temperate Forest Soils.” In Forest Growth Responses to the Pollution Climate of the 21st Century, edited by L. J. Sheppard and J. N. Cape, 89–98. Netherlands, Dordrecht: Springer.
  • Smith, W. H. 2012 Dec 6. Air Pollution and Forests: Interactions between Air Contaminants and Forest Ecosystems. In Springer Science & Business Media: US.
  • Sodhi, N. S., L. P. Koh, B. W. Brook, and P. K. Ng. “Southeast Asian Biodiversity: An Impending Disaster.” Trends in Ecology & Evolution 19(12): 654–660. 2004 Dec 1. doi:10.1016/j.tree.2004.09.006.
  • Song, H., K. Liu, and Y. N. Fan. 2016. “Analysis on spatial and temporal changes of forest degradation in Qinling Mountainous region,” Popul. Resour. Environ 5, 153–157.
  • Sonwani, S., and V. Maurya. 2018. “Impact of Air Pollution on the Environment and Economy.” Air Pollution: Sources, Impacts and Controls, 1:113.
  • Takahashi, M., Z. Feng, T. A. Mikhailova, O. V. Kalugina, O. V. Shergina, L. V. Afanasieva, R. K. Heng, N. M. Abd Majid, and H. Sase. 2020. “Air Pollution Monitoring and Tree and Forest Decline in East Asia: A Review.” Science of the Total Environment 18 (Jun): 140288. doi:10.1016/j.scitotenv.2020.140288.
  • Takamatsu, T., H. Sase, and J. Takada. “Some Physiological Properties of Cryptomeria Japonica Leaves from Kanto, Japan: Potential Factors Causing Tree Decline.” Canadian Journal of Forest Research 31(4): 663–672. 2001a Apr 1. doi:10.1139/x00-204.
  • Takamatsu, T., H. Sase, J. Takada, and R. Matsushita. 2001b. “Annual Changes in Some Physiological Properties of Cryptomeria Japonica Leaves from Kanto, Japan.” Water, Air, and Soil Pollution 130 (1, Aug): 941–946. doi:10.1023/A:1013962902375.
  • Tanaka, H., T. Hiyama, N. Kobayashi, H. Yabuki, Y. Ishii, R. V. Desyatkin, T. C. Maximov, and T. Ohta. 2008. “Energy Balance and Its Closure over a Young Larch Forest in Eastern Siberia Agric.” Forest Meteorol 148 (12): 1954–1967. doi:10.1016/j.agrformet.2008.05.006.
  • Tao, Z., and A. K. Jain. 2005. “Modeling of Global Biogenic Emissions for Key Indirect Greenhouse Gases and Their Response to Atmospheric CO 2 Increases and Changes in Land Cover and Climate.” J. Geophys. Res. Atmos 110 (D21): 1. doi:10.1029/2005JD005874.
  • Tasoglou, A., and S. N. Pandis. 2015. “Formation and Chemical Aging of Secondary Organic Aerosol during the Caryophyllene Oxidation.” Atmos. Chem. Phys 15 (11): 6035–6046. doi:10.5194/acp-15-6035-2015.
  • Taylor, K. E., R. J. Stouffer, and G. A. Meehl. 2012. “An Overview of CMIP5 and the Experiment Design.” Bulletin of the American Meteorological Society 93 (4): 485–498. doi:10.1175/BAMS-D-11-00094.1.
  • Thomley, J. H. M., and M. G. R. Cannell. 1996. “Forest Responses to Elevated [CO2], Temperature and Nitrogen Supply, Including Water Dynamics: Model Generated Hypotheses Compared with Observations.” Plant, Cell & Environment 19: 1331–1348.
  • Trivedi, L., and P. S. Dubey. 1983. Vegetation Damage Due to Satpura Thermal Power Station at North Betul Forest Division. InAir Pollution, Problems and Perspectives: Proceeding [S] of All India Seminar on Air Pollution, Indore, April 19-21, 1982/edited by DN Rao, TP Sharma, Udai Raj Singh; Foreworded by AN Varma 1983. Bhopal, India: Environmental Planning & Coordination Organization.
  • Ulziisaikhan, V. “Impact Assessment of Climate Change on Forest Ecosystem in Mongolia.” InRegional Workshop on Climate Change Vulnerability and Adaptation in Asia and the Pacific 1996 Jan 15, Manila, Philipinnes, (pp. 1–10).
  • UNECE. 2005. Europe’s Forests in a Changing Environment, 60. Geneva: Federal Research Centre for Forestry and Forest Products.
  • Vandemeulebroucke, I., S. Caluwaerts, and N. Van Den Bossche. 2021. “Factorial Study on the Impact of Climate Change on Freeze-Thaw Damage, Mould Growth and Wood Decay in Solid Masonry Walls in Brussels.” Buildings 11 (3, Mar): 134. doi:10.3390/buildings11030134.
  • Vij, B. M., L. Trivedi, A. Shevade, and P. S. Dubey. 1983. “Chlorophyll Damage in Tree Species Due to Air Pollution.” Biological Bulletin of India 3 (3): 193–194.
  • Vlasova, T. M., B. I. Kovalev, and A. N. Filipchuk. 1992. Effects of Point Source Atmospheric Pollution on boreal-forest Vegetation of Northwestern Siberia.Anchorage, AK, USA: National Park Service. No. AD-P=007320/5/XAB
  • Wagg, S., G. Mills, F. Hayes, S. Wilkinson, and W. J. Davies. 2013. “Stomata are Less Responsive to Environmental Stimuli in High Background Ozone in Dactylis Glomerata and Ranunculus Acris.” Environ Poll 175: 82–91. doi:10.1016/j.envpol.2012.11.027.
  • Wang, M. H., L. K. Wang, and N. K. Shammas. “Glossary of Climate Change, Global Warming and Ozone Layer Protection.” In HANDBOOK OF ENVIRONMENT AND WASTE MANAGEMENT: Acid Rain AND Greenhouse Gas Pollution Control. 2020. 689–718.
  • Wang, X. N., L. Qu, Q. Mao, M. Watanabe, Y. Hoshika, A. Koyama, K. Kawaguchi, Y. Tamai, and T. Koike. 2015. “Ectomycorrhizal Colonization and Growth of the Hybrid Larch F1 under Elevated CO2 and O3.” Environmental Pollution 197: 116–126. doi:10.1016/j.envpol.2014.11.031.
  • Wang, X. N., E. Agathokleous, L. Y. Qu, M. Watanabe, and T. Koike. 2016. “Effects of CO2 and O3 on the Interaction between Root of Woody Plants and Ectomycorrhizae.” Journal of Agriculture Meteorology 72 (2): 95–105. Open access doi:10.2480/agrmet.D-14-00045
  • Wittig, V. E., E. A. Ainsworth, S. L. Naidu, D. F. Karnosky, and S. P. Long. 2009. “Quantifying the Impact of Current and Future Tropospheric Ozone on Tree Biomass, Growth, Physiology and Biochemistry: A Quantitative Meta‐analysis.” Global Change Biology 15 (2, Feb): 396–424. doi:10.1111/j.1365-2486.2008.01774.x.
  • World Resource Institute (2017) Exploring Indonesia’s Long and Complicated History of Forest Fires February 16, 2017 by Andres Chamorro, Susan Minnemeyer and Sarah Sargent. https://www.wri.org/insights/exploring-indonesias-long-and-complicated-history-forest-fires
  • Wu, C., D. Chen, J. Shen, X. Sun, and S. Zhang. 2021b Feb 15. “Estimating the Distribution and Productivity Characters of Larix Kaempferi in Response to Climate Change.” Journal of Environmental Management 280: 111633. doi:10.1016/j.jenvman.2020.111633.
  • Wu, L., X. Ma, X. Dou, J. Zhu, and C. Zhao. 2021a Nov 20. “Impacts of Climate Change on Vegetation Phenology and Net Primary Productivity in Arid Central Asia.” Science of the Total Environment 796: 149055. doi:10.1016/j.scitotenv.2021.149055.
  • Xiang, W., M. Zhao, Z. Zhao, P. Lei, S. Ouyang, X. Deng, X. Zhou, and C. Peng. Simulating Carbon Sequestration Capacity of Forests in Subtropical Area: A Case Study in Hunan Province, Southern China. In L.Pawłowski, Z. Litwińczuk, G. Zhou, Editor.The Role of Agriculture in Climate Change Mitigation 2020. May 7, 95–105;CRC Press. London.
  • Yamaguchi, T., M. Watanabe, I. Noguchi, and T. Koike. 2017. Tree Decline at the Somma of Lake Mashu in Northern Japan. In T. Izuta, Editor. Air Pollution Impacts on Plants in East Asia Vol. 2017;135–150. Springer. Tokyo.
  • Yambe, Y. 1978. “Decline of Trees and Microbial Florae as the Index of Pollution in Some Urban areasBull.” Forestry and Forest Products Research Institute 301: 119–129.
  • Yim, S. H., Y. Gu, M. A. Shapiro, and B. Stephens. “Air Quality and Acid Deposition Impacts of Local Emissions and Transboundary Air Pollution in Japan and South Korea.” Atmospheric Chemistry and Physics 19(20): 13309–13323. 2019 Oct 28. doi:10.5194/acp-19-13309-2019.
  • Yoshitake, T., and H. Masuda. 1986. “Study on Unusual Defoliation of Pinus Strobus etc. in a Region of Tomakomai [Japan].” Bulletin of the Forestry and Forest Products Research Institute (Japan), 1:1-28.
  • Zafar, M. W., M. Shahbaz, F. Hou, and A. Sinha. 2019 Mar 1. “From Nonrenewable to Renewable Energy and Its Impact on Economic Growth: The Role of Research & Development Expenditures in Asia-Pacific Economic Cooperation Countries.” Journal of Cleaner Production 212: 1166–1178. doi:10.1016/j.jclepro.2018.12.081.
  • Zhang, W., Z. Feng, X. Wang, and J. Niu. 2012 Apr 1. “Responses of Native Broadleaved Woody Species to Elevated Ozone in Subtropical China.” Environmental Pollution 163: 149–157. doi:10.1016/j.envpol.2011.12.035.
  • Zhang, L., L. Hou, D. Guo, L. Li, and X. Xu. 2017. “Interactive Impacts of Nitrogen Input and Water Amendment on Growing Season Fluxes of CO2, CH4, and N2O in a Semiarid Grassland, Northern China.” Sci. Total Environ 578: 523–534. doi:10.1016/j.scitotenv.2016.10.219.
  • Zhou, S. S., T. Apk, S. H. Sun, M. Sadiq, C. L. Heald, and J. A. Geddes. 2018. “Coupling between Surface Ozone and Leaf Area Index in a Chemical Transport Model: Strength of Feedback and Implications for Ozone Air Quality and Vegetation Health.” Atmos Chem Phys 18 (19): 14133–14148. doi:10.5194/acp-18-14133-2018.
  • Zong, X., X. Tian, and Y. Yin. 2020. “Impacts of Climate Change on Wildfires in Central Asia.” Forests 11 (8, Aug): 802. doi:10.3390/f11080802.