Hostname: page-component-848d4c4894-pftt2 Total loading time: 0 Render date: 2024-06-03T16:44:07.431Z Has data issue: false hasContentIssue false

Research into land atmosphere interactions supports the sustainable development agenda

Published online by Cambridge University Press:  14 February 2024

Garry Hayman*
Affiliation:
Hydro-Climate Risks Science Area, UK Centre for Ecology & Hydrology, Wallingford, UK
Benjamin Poulter
Affiliation:
National Aeronautics and Space Administration, Goddard Space Flight Center, Greenbelt, MD, USA
Sachin D. Ghude
Affiliation:
Indian Institute of Tropical Meteorology, Ministry of Earth Science, Pune, India
Eleanor Blyth
Affiliation:
Hydro-Climate Risks Science Area, UK Centre for Ecology & Hydrology, Wallingford, UK
Vinayak Sinha
Affiliation:
Department of Earth and Environmental Sciences, Indian Institute of Science Education and Research Mohali, Mohali, Punjab, India
Sally Archibald
Affiliation:
School of Animal, Plant and Environmental Sciences, University of the Witwatersrand, Johannesburg, South Africa
Kirsti Ashworth
Affiliation:
Lancaster Environment Centre, Lancaster University, Lancaster, UK
Victoria Barlow
Affiliation:
Hydro-Climate Risks Science Area, UK Centre for Ecology & Hydrology, Wallingford, UK
Silvano Fares
Affiliation:
Institute for Agriculture and Forestry Systems in the Mediterranean, National Research Council of Italy, Naples, Italy
Gregor Feig
Affiliation:
South African Environmental Observation Network, Pretoria, South Africa Department of Geography, Geoinformatics & Meteorology, University of Pretoria, Pretoria, South Africa
Tetsuya Hiyama
Affiliation:
Institute for Space-Earth Environmental Research, Nagoya University, Nagoya, Japan
Jiming Jin
Affiliation:
College of Resources and Environment, Yangtze University, Hubei, China
Sirkku Juhola
Affiliation:
Faculty of Biological and Environmental Sciences, University of Helsinki, Helsinki, Finland
Meehye Lee
Affiliation:
Department of Earth and Environmental Sciences, Korea University, Seoul, South Korea
Sebastian Leuzinger
Affiliation:
School of Science, Auckland University of Technology, Auckland, New Zealand
Miguel D. Mahecha
Affiliation:
Institute for Earth System Science and Remote Sensing, Leipzig University, Leipzig, Germany RSC4Earth, Helmholtz Centre for Environmental Research, Leipzig, Germany
Xianhong Meng
Affiliation:
Northwest Institute of Eco-Environment and Resources, Chinese Academy of Sciences, Lanzhou, China
David Odee
Affiliation:
Kenya Forest Research Institute, Nairobi, Kenya, UK Centre for Ecology & Hydrology, Edinburgh, UK Biodiversity Science Area, UK Centre for Ecology & Hydrology, Edinburgh, UK
Gemma Purser
Affiliation:
Atmospheric Chemistry & Effects Science Area, UK Centre for Ecology & Hydrology, Edinburgh, UK
Hisashi Sato
Affiliation:
Research Institute for Global Change, Japan Agency for Marine-Earth Science and Technology, Yokohama, Kanagawa, Japan
Pallavi Saxena
Affiliation:
Department of Environmental Science, Hindu College, University of Delhi, Delhi, India
Valiyaveetil S. Semeena
Affiliation:
Hydro-Climate Risks Science Area, UK Centre for Ecology & Hydrology, Wallingford, UK
Allison Steiner
Affiliation:
Climate and Space Sciences and Engineering, University of Michigan, Ann Arbor, MI, USA
Xuemei Wang
Affiliation:
Institute for Environmental and Climate Research, Jinan University, Guangzhou, China
Stefan Wolff
Affiliation:
Multiphase Chemistry Department, Max Planck Institute for Chemistry, Mainz, Germany Max Planck Institute for Chemistry, INPA/Max Planck Project, Manaus, Brazil
*
Corresponding author: Garry Hayman; Email: garr@ceh.ac.uk

Abstract

Non-technical summary

Greenhouse gas emissions and land use change – from deforestation, forest degradation, and agricultural intensification – are contributing to climate change and biodiversity loss. Important land-based strategies such as planting trees or growing bioenergy crops (with carbon capture and storage) are needed to achieve the goals of the Paris Climate Agreement and to enhance biodiversity.

The integrated Land Ecosystems Atmospheric Processes Study (iLEAPS) is an international knowledge-exchange and capacity-building network, specializing in ecosystems and their role in controlling the exchange of water, energy and chemical compounds between the land surface and the atmosphere. We outline priority directions for land–atmosphere interaction research and its contribution to the sustainable development agenda.

Technical summary

Greenhouse-gas emissions from human activities and land use change (from deforestation, forest degradation, and agricultural intensification) are contributing to climate change and biodiversity loss. Afforestation, reforestation, or growing bioenergy crops (with carbon capture and storage) are important land-based strategies to achieve the goals of the Paris Climate Agreement and to enhance biodiversity. The effectiveness of these actions depends on terrestrial ecosystems and their role in controlling or moderating the exchange of water, heat, and chemical compounds between the land surface and the atmosphere.

The integrated Land Ecosystems Atmospheric Processes Study (iLEAPS), a global research network of Future Earth, enables the international community to communicate and remain up to date with developments and concepts about terrestrial ecosystems and their role in global water, energy, and biogeochemical cycles. Covering critically important topics such as fire, forestry, wetlands, methane emissions, urban areas, pollution, and climate change, the iLEAPS Global Research Programme sits center stage for some of the most important environmental questions facing humanity. In this paper, we outline the new challenges and opportunities for land–atmosphere interaction research and its role in supporting the broader sustainable development agenda.

Social Media Summary

Future directions for research into land–atmosphere interactions that supports the sustainable development agenda

Type
Intelligence Briefing
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2024. Published by Cambridge University Press

1. Introduction

The integrated Land Ecosystems Atmospheric Processes Study (iLEAPS) was formed in March 2004 to build an international community of practice to investigate the interactions between terrestrial ecosystems and the atmosphere. Originally part of the International Geosphere-Biosphere Programme, iLEAPS became a global research network of Future Earth in 2014.

The first decade of iLEAPS had an emphasis on creating new ways to observe and model the land–atmosphere continuum. Suni et al. (Reference Suni, Guenther, Hansson, Kulmala, Andreae, Arneth, Artaxo, Blyth, Brus, Ganzeveld, Kabat, de Noblet-Ducoudré, Reichstein, Reissell, Rosenfeld and Seneviratne2015) highlighted the iLEAPS contribution to the support and development of networks of long-term flux stations and large-scale land–atmosphere observation platforms. iLEAPS promoted the integration of data from remote sensing, ground-based observations, and other sources into data cubes (Mahecha et al., Reference Mahecha, Gans, Brandt, Christiansen, Cornell, Fomferra, Kraemer, Peters, Bodesheim, Camps-Valls, Donges, Dorigo, Estupinan-Suarez, Gutierrez-Velez, Gutwin, Jung, Londoño, Miralles, Papastefanou and Reichstein2020) and products (e.g. FLUXCOM (Jung et al., Reference Jung, Schwalm, Migliavacca, Walther, Camps-Valls, Koirala, Anthoni, Besnard, Bodesheim, Carvalhais, Chevallier, Gans, Goll, Haverd, Köhler, Ichii, Jain, Liu, Lombardozzi and Reichstein2020)). The understanding gained contributed to advances in land–surface models that represent the role of land cover changes and land–atmosphere feedback processes in the Earth system.

In this second decade, the focus has shifted to the human influence on these ecosystem–atmosphere interactions and the implications for resource use and sustainable development. While the impact on the natural environment is still investigated (He et al., Reference He, Clifton, Felker-Quinn, Fulgham, Juncosa Calahorrano, Lombardozzi, Purser, Riches, Schwantes, Tang, Poulter and Steiner2021), socio-economic aspects are also needed to cover mitigation and adaptation to climate change. The role of iLEAPS is ever more important in bringing together scientists to advance the knowledge of the complex Earth system and the people within it. Here we outline the new challenges and opportunities that motivate a new decadal iLEAPS roadmap for land–atmosphere interaction research.

2. The environmental challenges

The growing human population, its increasing demand for natural resources and the transformation and releases of materials, often harmful, to air, water, and land, is outstripping the capacity of the natural world to replenish or process them. Land cover changes due to deforestation, forest degradation, and agricultural intensification are major drivers of biodiversity loss (IPBES, 2019).

Climate change, caused by anthropogenic emissions of greenhouse gases (GHGs) and other compounds that change the Earth's radiative balance, is one of the many human pressures on the Earth system (IPCC, 2021). Higher global temperature results in greater impacts on ecosystems (Figure 1) and more extreme weather events, which are becoming more frequent and their impacts more severe. Their nature and consequences vary across the globe, with each region or country exposed to a different combination of hazards and risks, and each with different capacities to respond, mitigate or adapt.

Figure 1. Key risks to terrestrial and freshwater ecosystems from climate change (Figure 2.11 in IPCC AR6 WGII [IPCC, 2022a]).

iLEAPS provides information, understanding, and coordination of the science of this complex system. The iLEAPS focus on how interacting physical, chemical, and biological processes transport and transform energy and matter through the land–atmosphere interface is critical in understanding the processes and impacts of climate and other planetary changes.

3. The policy and scientific response

The Paris Agreement of the UN Framework Convention on Climate Change (UNFCCC, 2015) signaled a change of focus from reducing emissions of GHG and near-term climate forcers to ‘holding the increase in global average temperature to well below 2 °C and pursuing efforts to limit the increase to 1.5 °C’. Meeting this goal requires immediate and sustained reductions in emissions, combined with CO2 removal, in which land-based measures – for example, ecosystem rehabilitation, afforestation (in appropriate systems), bioenergy, the latter combined with carbon capture and storage – play a pivotal role (IPCC, 2018). To inform the UNFCCC process, the Intergovernmental Panel on Climate Change (IPCC) has produced a series of assessment reports on climate change (e.g. IPCC (2021, 2022a, 2022b)) and topical special reports (e.g. IPCC (2018, 2019a, 2019b)), which highlight knowledge gaps and the need for improved information, for example, about glaciers, fire, and methane sources.

The UN Agenda for Sustainable Development has developed 17 Sustainable Development Goals (SDGs), which recognize that improving human life and reducing inequality must go hand-in-hand with tackling climate change and working to preserve oceans and forests. The UN Convention on Biological Diversity, inspired by the growing commitment to sustainable development, is another significant milestone. The complex and multiple connections between climate and biodiversity have been recognized in the international policy arena (Pörtner et al., Reference Pörtner, Scholes, Agard, Archer, Arneth, Bai, Barnes, Burrows, Chan, Cheung, Diamond, Donatti, Duarte, Eisenhauer, Foden, Gasalla, Handa, Hickler, Hoegh-Guldberg and Ngo2021). There is also increased interest in nature-based solutions, defined by the International Union for Conservation of Nature as actions to protect, sustainably manage, and restore ecosystems that address societal challenges.

In response, the international research community established Future Earth to bridge the historical divide between scientific knowledge and societal action by advocating research that supports global sustainability (van der Hel, Reference van der Hel2016). Future Earth has created new structures (Suni et al., Reference Suni, Juhola, Korhonen-Kurki, Käyhkö, Soini and Kulmala2016), in which global research networks such as iLEAPS play an important role, in identifying and addressing key scientific gaps.

4. ‘iLEAPS' facilitation

The iLEAPS science plan (included as Supplementary Material) focuses on three overlapping terrestrial systems: (a) natural, (b) managed land, and (c) urban, together with three critical cross-cutting themes: (d) cold regions (e) arid/semi-arid regions, and (f) wetlands. These are indicated schematically in Figure 2, which also shows the key land–atmosphere processes of these focal systems and themes. In the following sections, we identify the environmental problems that iLEAPS is currently addressing and where its efforts will be focused in the future.

  1. (a) Natural ecosystems

Figure 2. Schematic of the three focal systems of interest to iLEAPS, their overlap, the three key cross-cutting themes, and the land–atmosphere processes involved.

More than 95% of the Earth's land surface is directly affected by human activities (Plumptre et al., Reference Plumptre, Baisero, Belote, Vázquez-Domínguez, Faurby, Jȩdrzejewski, Kiara, Kühl, Benítez-López, Luna-Aranguré, Voigt, Wich, Wint, Gallego-Zamorano and Boyd2021). Natural ecosystems, defined as land that has not been actively altered or managed for at least a generation, play a vital role in the Earth system. Forests have the greatest impact on climate, both directly through albedo and surface roughness, and the exchange of heat, energy and trace compounds, and indirectly via carbon storage. Forests also provide a vital range of ecosystem services, including habitat for biodiversity, food, and fiber for people. Forests cover 31% of the global land area but are not distributed uniformly (FAO, 2022). Their loss has serious implications for climate and ecosystem services. Natural grass systems also provide essential ecosystem services (e.g. below ground carbon sequestration (Dass et al., Reference Dass, Houlton, Wang and Warlind2018; Retallack, Reference Retallack2013; Ryan et al., Reference Ryan, Williams and Grace2011)), provide livelihoods for a significant proportion of the human population and support a unique and adapted biodiversity.

A study of historic losses of carbon due to land use change, mainly deforestation, suggests that global biomass could potentially be ~400 Pg carbon greater than at present (Erb et al., Reference Erb, Kastner, Plutzar, Bais, Carvalhais, Fetzel, Gingrich, Haberl, Lauk, Niedertscheider, Pongratz, Thurner and Luyssaert2018), equivalent to ~40 years of CO2 emissions at current levels. While this is an important rationale for promoting tree planting as a climate mitigation option (e.g. the Bonn Challenge, FAO (2022)), such solutions also need to balance the diverse ecosystem services that forests provide (Lewis et al., Reference Lewis, Mitchard, Prentice, Maslin and Poulter2019) or to avoid the unintended consequences of such actions on the ecosystem service provision in grassland systems (Bond et al., Reference Bond, Stevens, Midgley and Lehmann2019). Ryan et al. (Reference Ryan, Williams and Grace2011) estimate that the savannah regions of southern Africa store between 18 and 24 Pg of carbon (split evenly between the soil and woody vegetation), which is of a similar magnitude that stored in the Congo Basin rainforests (30 Pg C).

The iLEAPS community is at the forefront of investigating the many trade-offs between carbon storage and other ecosystem services and highlighting the consequences of tree species changes due to climate change as well as human deforestation and afforestation activities. In particular, iLEAPS scientists have emphasized the need to consider the interactions between vegetation, local microclimate, fire occurrence, air quality, and human health through the emissions of biogenic volatile organic compounds and bio-aerosols such as pollen or fungal spores.

  1. (b) Managed land

Managed land refers to that cultivated for agricultural food crops, for agroforestry use, silviculture, plantations and pastures grown for biomass for energy, timber, industrial products (e.g. paper, rubber), and livestock and may include management interventions, harvesting, thinning, the use of fire, and application of soil improvers, for example, agrochemical nitrogen and phosphorous fertilizers or natural fertilizers (Ogle et al., Reference Ogle, Domke, Kurz, Rocha, Huffman, Swan, Smith, Woodall and Krug2018). Managed lands collectively represent one of the most dynamically changing components of the land–biosphere–atmosphere system, as they keep pace with the increased demands of food, shelter, and energy for the world's rapidly increasing population. Land management has direct impacts on carbon stocks, air quality (e.g. contribution to aerosols), and is associated with a range of GHG emissions and also the removal of such gases through uptake and deposition.

The global tree count census (Crowther et al., Reference Crowther, Glick, Covey, Bettigole, Maynard, Thomas, Smith, Hintler, Duguid, Amatulli, Tuanmu, Jetz, Salas, Stam, Piotto, Tavani, Green, Bruce, Williams and Bradford2015) neglected the presence of cropland trees across the globe and Bastin et al. (Reference Bastin, Finegold, Garcia, Mollicone, Rezende, Routh, Zohner and Crowther2019) excluded croplands while proposing reforestation as a tool for carbon sequestration. There is at least 45 million ha of agroforestry land, which is expected to expand with the ongoing tree planting initiatives in degraded land (FAO, 2022). Two-thirds of the chemistry-climate models used by the IPCC exclude cropland trees in their land-use land-cover module (Mishra et al., Reference Mishra, Sinha, Kumar, Barth, Hakkim, Kumar, Kumar, Datta, Guenther and Sinha2021). The ability to simulate dynamic land-use changes related to biomass production or agricultural crop rotation has yet to be undertaken by such chemistry-climate models and is currently reflected in site-based or regional-scale process models (Havermann et al., Reference Havermann, Ghirardo, Schnitzler, Nendel, Hoffmann, Kraus and Grote2022).

The impact of air pollution, such as ozone, acid deposition, and particulate matter, on agricultural crops has been assessed in modeling studies at the global scale (Van Dingenen et al., Reference Van Dingenen, Dentener, Raes, Krol, Emberson and Cofala2009). An important finding of the Tropospheric Ozone Assessment Report (Mills et al., [Reference Mills, Pleijel, Malley, Sinha, Cooper, Schultz, Neufeld, Simpson, Sharps, Feng, Gerosa, Harmens, Kobayashi, Saxena, Paoletti, Sinha and Xu2018]) is that some regions, in particular Africa and South America, have very limited air pollution monitoring, making a complete global assessment difficult. The importance of understanding these atmosphere–biosphere feedbacks is very relevant to the UN sustainable development goal on zero hunger.

  1. (c) Urban

The majority of the world's population now lives within an urban environment. The combined effect of global climate change and rapid urban growth, in tandem with economic and industrial development, will induce or exacerbate a number of the urban environmental problems (Figure 3).

Figure 3. Schematic diagram of the impacts of urbanization on climate and air quality (adapted from Wang et al. [Reference Wang, Wu, Zhang, Cohen, Pang, Bouarar, Wang and Brasseur2017]).

Green infrastructure and other nature-based solutions are increasingly considered to provide co-benefits for urban areas. They can have benefits for carbon sequestration and adaptation to climate risks and to mediate air quality and heat (Grote et al., Reference Grote, Samson, Alonso, Amorim, Cariñanos, Churkina, Fares, Thiec, Niinemets, Mikkelsen, Paoletti, Tiwary and Calfapietra2016). However, in urban areas, there can also be trade-offs between adaptation and mitigation (Landauer et al., Reference Landauer, Juhola and Söderholm2015; Locatelli et al., Reference Locatelli, Pavageau, Pramova and Di Gregorio2015). The emissions of volatile organic compounds from urban vegetation in the presence of high NOx sources may increase ozone and secondary aerosol formation in the urban atmosphere.

The number of premature deaths from exposure to outdoor air pollution is projected to increase from ~3 million people globally in 2010 to 6–9 million in 2060. The distribution of these premature deaths across the globe is unequal, with the highest number of deaths in China and India (OECD, 2016). In addition, evidence suggests that air pollution may be linked to a decrease in life satisfaction, and an increase in associated negative mental health outcomes and therefore may have wider reaching impacts, both societally and economically (Lu, Reference Lu2020).

  1. (d) Cold/high-elevation regions

The cryosphere is especially sensitive and changes could significantly impact on the natural environment and human society. The Arctic is warming faster than the global average (IPCC, 2021), causing perturbations to the terrestrial water and carbon cycles in this region. Warming may have already shifted some ecosystems from net carbon sinks toward carbon-neutral or carbon sources, although it remains a challenge to determine the net ecosystem response across the circumpolar scale (Schuur et al., Reference Schuur, Abbott, Commane, Ernakovich, Euskirchen, Hugelius, Grosse, Jones, Koven, Leshyk, Lawrence, Loranty, Mauritz, Olefeldt, Natali, Rodenhizer, Salmon, Schädel, Strauss and Turetsky2022). The interactions and feedbacks to the atmosphere, regional climate, and water resources, make quantitative forecasts challenging.

Understanding the consequences of mountain glacier retreat is vital, since glaciers store and supply fresh water to lowland areas (Meyer et al., Reference Meyer, Strayer, Wallace, Eggert, Helfman and Leonard2007). In addition, plant productivity is generally limited by low temperatures, short growing seasons, and aridity (Paquette & Hargreaves, Reference Paquette and Hargreaves2021). Current warming trends and decreasing precipitation in continental interiors would change the boundaries of these limitations, resulting in shifts of species (Gauthier et al., Reference Gauthier, Bernier, Burton, Edwards, Isaac, Isabel, Jayen, Le Goff and Nelson2014).

Warming can also have consequences for the permafrost in these regions, such that permafrost carbon, which is equivalent to about 40% (1,460– 1,600 Pg C, Schuur et al. (Reference Schuur, Abbott, Commane, Ernakovich, Euskirchen, Hugelius, Grosse, Jones, Koven, Leshyk, Lawrence, Loranty, Mauritz, Olefeldt, Natali, Rodenhizer, Salmon, Schädel, Strauss and Turetsky2022)) of total global carbon within soils and biomass (Friedlingstein et al. (Reference Friedlingstein, O'Sullivan, Jones, Andrew, Gregor, Hauck, Le Quéré, Luijkx, Olsen, Peters, Peters, Pongratz, Schwingshackl, Sitch, Canadell, Ciais, Jackson, Alin, Alkama and Zheng2022): permafrost =  1,400, soils = 1,700, vegetation = 450 Pg C) is projected to decrease (IPCC, 2021). Finally, wildfire, one of the most significant disturbance agents at high latitudes, is expected to increase in frequency and severity (UNEP, Reference Sullivan, Baker and Kurvits2022), exacerbating changes in these sensitive regions.

  1. (e) Arid/semi-arid regions

Semi-arid regions are geographically located between the arid and humid regions, where land–atmosphere interactions are stronger because of abrupt change between moisture regimes. This makes them simultaneously more sensitive to climate change and more influential in terms of feedbacks to the global carbon and hydrological cycles (Ahlström et al., Reference Ahlström, Raupach, Schurgers, Smith, Arneth, Jung, Reichstein, Canadell, Friedlingstein, Jain, Kato, Poulter, Sitch, Stocker, Viovy, Wang, Wiltshire, Zaehle and Zeng2015; Poulter et al., Reference Poulter, Frank, Ciais, Myneni, Andela, Bi, Broquet, Canadell, Chevallier, Liu, Running, Sitch and van der Werf2014). They also exert powerful regional influences. Thus, the severe drought experienced across the center of North China has been attributable to stronger sensible heating in the western arid regions in addition to direct climate warming (Huang et al., Reference Huang, Zhou, Chen, Zhou, Wei, Zhang, Gao, Wei and Hou2013; Liu et al., Reference Liu, Li, Huang, Zhu and Wang2019).

The structure and carbon budget of semi-arid vegetation are also under the strong control of wildfire. Approximately 3% of the global land surface burns annually, which represents a significant but poorly understood mechanism for the exchange of energy and matter between the land surface and the atmosphere, and longer-term alterations to the characteristics of the land surface (Archibald et al., Reference Archibald, Lehmann, Belcher, Bond, Bradstock, Daniau, Dexter, Forrestel, Greve, He, Higgins, Hoffmann, Lamont, McGlinn, Moncrieff, Osborne, Pausas, Price, Ripley and Zanne2018). Earth system models examining the impact of altered fire regimes indicate the potential for significant increases in global mean surface air temperature, decreased net radiation, and latent heat (Li et al., Reference Li, Lawrence and Bond-Lamberty2017).

The geographical distributions, frequency, and intensity of wildfires are projected to change under current warming trends. A comparison of global ‘fire-on’ and ‘fire-off’ simulations shows that wildfire maintains vast areas of humid C4 grasslands and savannahs, especially in South America and Africa, against its climate potential to form forest (Bond et al., Reference Bond, Woodward and Midgley2005). Meta-analysis using data from savannahs across the world indicates that vegetation–fire–climate relationships differ across continents (Lehmann et al., Reference Lehmann, Anderson, Sankaran, Higgins, Archibald, Hoffmann, Hanan, Williams, Fensham, Felfili, Hutley, Ratnam, San Jose, Montes, Franklin, Russell-Smith, Ryan, Durigan, Hiernaux and Bond2014), but observational studies in Africa suggest that the existence of savannah ecosystems there requires intensive disturbances (fire, herbivory) (Sankaran et al., Reference Sankaran, Hanan, Scholes, Ratnam, Augustine, Cade, Gignoux, Higgins, Le Roux, Ludwig, Ardo, Banyikwa, Bronn, Bucini, Caylor, Coughenour, Diouf, Ekaya, Feral and Zambatis2005).

  1. (f) Wetlands

Wetlands are ecosystems in which mineral or peat soils are water saturated or where surface inundation dominates the soil biogeochemistry and determines the ecosystem species composition (USEPA, 2010). They are concentrated in two broad latitudinal bands: one rich in peatlands that spans the boreal and subarctic zones and a second covering the tropics and sub-tropics that contain vast swamps and seasonally inundated floodplains (Kirschke et al., Reference Kirschke, Bousquet, Ciais, Saunois, Canadell, Dlugokencky, Bergamaschi, Bergmann, Blake, Bruhwiler, Cameron-Smith, Castaldi, Chevallier, Feng, Fraser, Heimann, Hodson, Houweling, Josse and Zeng2013). Wetlands are an important component of the global water and carbon cycles, influencing groundwater balance, and river flow (Melton et al., Reference Melton, Wania, Hodson, Poulter, Ringeval, Spahni, Bohn, Avis, Beerling, Chen, Eliseev, Denisov, Hopcroft, Lettenmaier, Riley, Singarayer, Subin, Tian, Zürcher and Kaplan2013), and collectively represent the largest natural source of methane (Saunois et al., Reference Saunois, Stavert, Poulter, Bousquet, Canadell, Jackson, Raymond, Dlugokencky, Houweling, Patra, Ciais, Arora, Bastviken, Bergamaschi, Blake, Brailsford, Bruhwiler, Carlson, Carrol and Zhuang2020). Boreal and subarctic wetlands store most of the global wetland soil carbon stock (Turetsky et al., Reference Turetsky, Kotowska, Bubier, Dise, Crill, Hornibrook, Minkkinen, Moore, Myers-Smith, Nykänen, Olefeldt, Rinne, Saarnio, Shurpali, Tuittila, Waddington, White, Wickland and Wilmking2014). In the tropics, trees subjected to permanent or periodic inundation have developed adaptive features to enhance oxygenation of their root systems, which facilitate the natural release of soil CH4 to the atmosphere (Pangala et al., Reference Pangala, Enrich-Prast, Basso, Peixoto, Bastviken, Hornibrook, Gatti, Marotta, Calazans, Sakuragui, Bastos, Malm, Gloor, Miller and Gauci2017).

Methane from wetlands, especially from the tropical wetlands, has been identified as a key driver of the increased concentrations of atmospheric methane and the shift in its isotopic composition (Oh et al., Reference Oh, Zhuang, Welp, Liu, Lan, Basu, Dlugokencky, Bruhwiler, Miller, Michel, Schwietzke, Tans, Ciais and Chanton2022). Increased wetland methane emissions will offset the climate benefits of any reductions in anthropogenic methane emissions (e.g. through the global methane pledge) (Comyn-Platt et al., Reference Comyn-Platt, Hayman, Huntingford, Chadburn, Burke, Harper, Collins, Webber, Powell, Cox, Gedney and Sitch2018; Zhang et al., Reference Zhang, Poulter, Feldman, Ying, Ciais, Peng and Li2023).

Substantial GHG emissions are released from forested peatlands in Southeast Asia that are being drained and replaced with perennial crops, such as oil-palm and pulpwood plantations (IPCC, 2019a). Restoring tropical peatlands has benefits not only for mitigating climate change but also for reducing fire risk and for biodiversity (Tan et al., Reference Tan, Carrasco, Sutikno and Taylor2022).

5. Future focus of iLEAPS and vision for 2035

The iLEAPS science plan (included as Supplementary Material) aims to provide a vision for the next decade and, for example, synergies with the SDG's, responses to the latest IPCC and IPBES assessment reports, and areas for further research. The science plan takes into account how research networks have matured over the past decade, as well as the emergence of novel measurement and monitoring from towers, aircraft, and space to inform model development, and the need to consider ‘big data’ approaches and data equity in addressing science questions related to climate change, climate mitigation and adaptation. Over the next decade, the science and applications vision for iLEAPS includes the following cross-cutting themes that advance the focal systems of interest in Figure 2:

6. Concluding remarks

As the planet experiences increasingly large-scale changes in atmospheric temperature, precipitation, and chemical composition, it is urgent that we understand the complex interactions of these changes with the land system to realize their full impact. Land–atmosphere interactions are central to a wide-ranging body of scientific enquiry, bringing vital understanding of small-scale processes (e.g. to create a healthier urban environment) through to managing large-scale landscapes (e.g. to unlock its climate mitigation potential) while maintaining essential ecosystem services. Any proposed changes to land use require us to understand the impact of atmospheric chemistry and meteorology on the functioning of the land-system.

With specialists and science leaders from across the world, and with expertise across the broad range of science covered by iLEAPS, this inclusive hub enables the international community to communicate and remain up to date with developments and concepts on this link in the earth-system chain. Covering critically important processes such as fire, forestry, wetlands, methane, urban areas, pollution, and climate change, it is evident that iLEAPS sits center stage of some of the most important and challenging environmental questions facing humanity.

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1017/sus.2024.3.

Acknowledgements

The iLEAPS Scientific Steering Committee (SSC) acknowledges the support of Future Earth to enable annual meetings of the SSC, which facilitated the preparation of this paper. The iLEAPS International Project Office (IPO) is grateful to the UK Centre for Ecology & Hydrology for hosting the IPO.

Author contributions

E.B., G.H., V.S., and B.P. conceived the idea of an iLEAPS paper. G.H. lead the preparation of the paper and all authors contributed to the writing of specific sections or to the review of the paper.

Funding statement

The iLEAPS Scientific Steering Committee acknowledges the grants received from Future Earth to cover the annual SSC meetings. Past and present members of the iLEAPS International Project Office (G.H., E.B., V.B., and V.S.S.) acknowledge the support of the UK Natural Environment Research Council (NERC), through: (a) its International Opportunities Fund, grant NE/P008615/1 and (b) the NC-International programme [NE/X006247/1] delivering National Capability. G.P. also acknowledges the support of the NERC International programme to UKCEH. K.A. is grateful to the Royal Society of London for their funding of her Dorothy Hodgkin Research Fellowship (DH150070).

Competing interest

The authors declare no conflict of interests.

Research transparency and reproducibility

No unpublished data or software has been used in this manuscript.

Figure 1 is taken from the IPCC AR6 WGII report. The IPCC allows reproduction of a limited number of figures or short excerpts of IPCC material free of charge and without formal written permission, provided that the material is not altered and the original source is properly acknowledged.

References

Ahlström, A., Raupach, M. R., Schurgers, G., Smith, B., Arneth, A., Jung, M., Reichstein, M., Canadell, J. G., Friedlingstein, P., Jain, A. K., Kato, E., Poulter, B., Sitch, S., Stocker, B. D., Viovy, N., Wang, Y. P., Wiltshire, A., Zaehle, S., & Zeng, N. (2015). The dominant role of semi-arid ecosystems in the trend and variability of the land CO2 sink. Science (New York, N.Y.), 348, 895899. https://doi.org/10.1126/science.aaa1668CrossRefGoogle ScholarPubMed
Archibald, S., Lehmann, C. E. R., Belcher, C. M., Bond, W. J., Bradstock, R. A., Daniau, A. L., Dexter, K. G., Forrestel, E. J., Greve, M., He, T., Higgins, S. I., Hoffmann, W. A., Lamont, B. B., McGlinn, D. J., Moncrieff, G. R., Osborne, C. P., Pausas, J. G., Price, O., Ripley, B. S., … Zanne, A. E. (2018). Biological and geophysical feedbacks with fire in the Earth system. Environmental Research Letters, 13, 033003. https://doi.org/10.1088/1748-9326/aa9eadCrossRefGoogle Scholar
Bastin, J.-F., Finegold, Y., Garcia, C., Mollicone, D., Rezende, M., Routh, D., Zohner, C. M., & Crowther, T. W. (2019). The global tree restoration potential. Science (New York, N.Y.), 365, 76. https://doi.org/10.1126/science.aax0848CrossRefGoogle ScholarPubMed
Bond, W. J., Stevens, N., Midgley, G. F., & Lehmann, C. E. R. (2019). The trouble with trees: Afforestation plans for Africa. Trends in Ecology & Evolution, 34, 963965. https://doi.org/10.1016/j.tree.2019.08.003CrossRefGoogle ScholarPubMed
Bond, W. J., Woodward, F. I., & Midgley, G. F. (2005). The global distribution of ecosystems in a world without fire. New Phytologist, 165, 525538. https://doi.org/10.1111/j.1469-8137.2004.01252.xCrossRefGoogle Scholar
Comyn-Platt, E., Hayman, G., Huntingford, C., Chadburn, S. E., Burke, E. J., Harper, A. B., Collins, W. J., Webber, C. P., Powell, T., Cox, P. M., Gedney, N., & Sitch, S. (2018). Carbon budgets for 1.5 and 2 °C targets lowered by natural wetland and permafrost feedbacks. Nature Geoscience, 11, 568573. https://doi.org/10.1038/s41561-018-0174-9CrossRefGoogle Scholar
Conradie, E. H., Van Zyl, P. G., Pienaar, J. J., Beukes, J. P., Galy-Lacaux, C., Venter, A. D., & Mkhatshwa, G. V. (2016). The chemical composition and fluxes of atmospheric wet deposition at four sites in South Africa. Atmospheric Environment, 146, 113131. https://doi.org/10.1016/j.atmosenv.2016.07.033CrossRefGoogle Scholar
Crowther, T. W., Glick, H. B., Covey, K. R., Bettigole, C., Maynard, D. S., Thomas, S. M., Smith, J. R., Hintler, G., Duguid, M. C., Amatulli, G., Tuanmu, M. N., Jetz, W., Salas, C., Stam, C., Piotto, D., Tavani, R., Green, S., Bruce, G., Williams, S. J., … Bradford, M. A. (2015). Mapping tree density at a global scale. Nature, 525, 201205. https://doi.org/10.1038/nature14967CrossRefGoogle ScholarPubMed
Dass, P., Houlton, B. Z., Wang, Y., & Warlind, D. (2018). Grasslands may be more reliable carbon sinks than forests in California. Environmental Research Letters, 13, 074027. https://doi.org/10.1088/1748-9326/aacb39CrossRefGoogle Scholar
Erb, K.-H., Kastner, T., Plutzar, C., Bais, A. L. S., Carvalhais, N., Fetzel, T., Gingrich, S., Haberl, H., Lauk, C., Niedertscheider, M., Pongratz, J., Thurner, M., & Luyssaert, S. (2018). Unexpectedly large impact of forest management and grazing on global vegetation biomass. Nature, 553, 7376. https://doi.org/10.1038/nature25138CrossRefGoogle ScholarPubMed
FAO. (2022). The state of the world's forests 2022. Forest pathways for green recovery and building inclusive, resilient and sustainable economies. The Food and Agriculture Organization. https://doi.org/10.4060/cb9360enGoogle Scholar
Friedlingstein, P., O'Sullivan, M., Jones, M. W., Andrew, R. M., Gregor, L., Hauck, J., Le Quéré, C., Luijkx, I. T., Olsen, A., Peters, G. P., Peters, W., Pongratz, J., Schwingshackl, C., Sitch, S., Canadell, J. G., Ciais, P., Jackson, R. B., Alin, S. R., Alkama, R., … Zheng, B. (2022). Global carbon budget 2022. Earth System Science Data, 14, 48114900. https://doi.org/10.5194/essd-14-4811-2022CrossRefGoogle Scholar
Gauthier, S., Bernier, P., Burton, P. J., Edwards, J., Isaac, K., Isabel, N., Jayen, K., Le Goff, H., & Nelson, E. A. (2014). Climate change vulnerability and adaptation in the managed Canadian boreal forest. Environmental Reviews, 22, 256285. https://doi.org/10.1139/er-2013-0064CrossRefGoogle Scholar
Grote, R., Samson, R., Alonso, R., Amorim, J. H., Cariñanos, P., Churkina, G., Fares, S., Thiec, D. L., Niinemets, Ü, Mikkelsen, T. N., Paoletti, E., Tiwary, A., & Calfapietra, C. (2016). Functional traits of urban trees: Air pollution mitigation potential. Frontiers in Ecology and the Environment, 14, 543550. https://doi.org/10.1002/fee.1426CrossRefGoogle Scholar
Hancock, S., Armston, J., Hofton, M., Sun, X., Tang, H., Duncanson, L. I., Kellner, J. R., & Dubayah, R. (2019). The GEDI simulator: A large-footprint waveform lidar simulator for calibration and validation of spaceborne missions. Earth and Space Science, 6, 294310. https://doi.org/10.1029/2018ea000506CrossRefGoogle ScholarPubMed
Havermann, F., Ghirardo, A., Schnitzler, J.-P., Nendel, C., Hoffmann, M., Kraus, D., & Grote, R. (2022). Modeling intra- and interannual variability of BVOC emissions from maize. Oil-Seed Rape, and Ryegrass. Journal of Advances in Modeling Earth Systems, 14, e2021MS002683. https://doi.org/10.1029/2021MS002683CrossRefGoogle Scholar
He, C., Clifton, O., Felker-Quinn, E., Fulgham, S. R., Juncosa Calahorrano, J. F., Lombardozzi, D., Purser, G., Riches, M., Schwantes, R., Tang, W., Poulter, B., & Steiner, A. L. (2021). Interactions between air pollution and terrestrial ecosystems: Perspectives on challenges and future directions. Bulletin of the American Meteorological Society, 102, E525E538. https://doi.org/10.1175/BAMS-D-20-0066.1CrossRefGoogle Scholar
Hernando, A., Puerto, L., Mola-Yudego, B., Manzanera, J., Garcia-Abril, A., Maltamo, M., & Valbuena, R. (2019). Estimation of forest biomass components using airborne LiDAR and multispectral sensors. iForest – Biogeosciences and Forestry, 12, 207213. https://doi.org/10.3832/ifor2735-012CrossRefGoogle Scholar
Huang, R., Zhou, D., Chen, W., Zhou, L., Wei, Z., Zhang, Q., Gao, X., Wei, G., & Hou, X. (2013). Recent advances in research on land–air interactions and their impact on climate in arid regions of northwestern China. Chinese Journal of Atmospheric Sciences, 37, 189210. https://doi.org/10.3878/j.issn.1006-9895.2012.12303Google Scholar
IPBES. (2019). Global assessment report on biodiversity and ecosystem services of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services. E. S. Brondizio, J. Settele, S. Díaz, and H. T. Ngo (editors). IPBES secretariat, Bonn, Germany. 1148 pages. https://doi.org/10.5281/zenodo.3831673CrossRefGoogle Scholar
IPCC. (2018). Global Warming of 1.5 °C, IPCC special report on the impacts of global warming of 1.5 °C above pre-industrial levels and related global greenhouse gas emission pathways, in the context of strengthening the global response to the threat of climate change, sustainable development, and efforts to eradicate poverty [Masson-Delmotte, V., P. Zhai, H.-O. Pörtner, D. Roberts, J. Skea, P.R. Shukla, A. Pirani, W. Moufouma-Okia, C. Péan, R. Pidcock, S. Connors, J.B.R. Matthews, Y. Chen, X. Zhou, M.I. Gomis, E. Lonnoy, T. Maycock, M. Tignor, and T. Waterfield (eds.)]. Cambridge University Press, Cambridge, UK and New York, NY, USA, 616 pp. https://doi.org/10.1017/9781009157940CrossRefGoogle Scholar
IPCC. (2019a). Climate Change and Land: an IPCC special report on climate change, desertification, land degradation, sustainable land management, food security, and greenhouse gas fluxes in terrestrial ecosystems [P.R. Shukla, J. Skea, E. Calvo Buendia, V. Masson-Delmotte, H.-O. Pörtner, D. C. Roberts, P. Zhai, R. Slade, S. Connors, R. van Diemen, M. Ferrat, E. Haughey, S. Luz, S. Neogi, M. Pathak, J. Petzold, J. Portugal Pereira, P. Vyas, E. Huntley, K. Kissick, M. Belkacemi, J. Malley, (eds.)]. Cambridge University Press, Cambridge, UK and New York, NY, USA, 896 pp. https://doi.org/10.1017/9781009157988CrossRefGoogle Scholar
IPCC. (2019b). IPCC Special Report on the Ocean and Cryosphere in a Changing Climate [H.-O. Pörtner, D.C. Roberts, V. Masson-Delmotte, P. Zhai, M. Tignor, E. Poloczanska, K. Mintenbeck, A. Alegría, M. Nicolai, A. Okem, J. Petzold, B. Rama, N.M. Weyer (eds.)]. Cambridge University Press, Cambridge, UK and New York, NY, USA, 755 pp. https://doi.org/https://doi.org/10.1017/9781009157964CrossRefGoogle Scholar
IPCC. (2021). Climate Change 2021: The Physical Science Basis. Working Group I contribution to the IPCC Sixth Assessment Report [Masson-Delmotte, V., P. Zhai, A. Pirani, S.L. Connors, C. Péan, S. Berger, N. Caud, Y. Chen, L. Goldfarb, M.I. Gomis, M. Huang, K. Leitzell, E. Lonnoy, J.B.R. Matthews, T.K. Maycock, T. Waterfield, O. Yelekçi, R. Yu, and B. Zhou (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 2391 pp. https://doi.org/10.1017/9781009157896CrossRefGoogle Scholar
IPCC. (2022a). Climate Change 2022: Impacts, Adaptation and Vulnerability. The Working Group II contribution to the IPCC Sixth Assessment Report. [H.-O. Pörtner, D.C. Roberts, M. Tignor, E.S. Poloczanska, K. Mintenbeck, A. Alegría, M. Craig, S. Langsdorf, S. Löschke, V. Möller, A. Okem, B. Rama (eds.)]. Cambridge University Press. Cambridge University Press, Cambridge, UK and New York, NY, USA, 3056 pp. https://doi.org/10.1017/9781009325844CrossRefGoogle Scholar
IPCC. (2022b). Climate Change 2022: Mitigation of Climate Change. The Working Group III contribution to the IPCC's Sixth Assessment Report. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA. https://doi.org/10.1017/9781009157926CrossRefGoogle Scholar
Jung, M., Koirala, S., Weber, U., Ichii, K., Gans, F., Camps-Valls, G., Papale, D., Schwalm, C., Tramontana, G., & Reichstein, M. (2019). The FLUXCOM ensemble of global land-atmosphere energy fluxes. Scientific Data, 6, 74. https://doi.org/10.1038/s41597-019-0076-8CrossRefGoogle ScholarPubMed
Jung, M., Schwalm, C., Migliavacca, M., Walther, S., Camps-Valls, G., Koirala, S., Anthoni, P., Besnard, S., Bodesheim, P., Carvalhais, N., Chevallier, F., Gans, F., Goll, D. S., Haverd, V., Köhler, P., Ichii, K., Jain, A. K., Liu, J., Lombardozzi, D., … Reichstein, M. (2020). Scaling carbon fluxes from eddy covariance sites to globe: Synthesis and evaluation of the FLUXCOM approach. Biogeosciences (Online), 17, 13431365. https://doi.org/10.5194/bg-17-1343-2020CrossRefGoogle Scholar
Kirschke, S., Bousquet, P., Ciais, P., Saunois, M., Canadell, J. G., Dlugokencky, E. J., Bergamaschi, P., Bergmann, D., Blake, D. R., Bruhwiler, L., Cameron-Smith, P., Castaldi, S., Chevallier, F., Feng, L., Fraser, A., Heimann, M., Hodson, E. L., Houweling, S., Josse, B., … Zeng, G. (2013). Three decades of global methane sources and sinks. Nature Geoscience, 6, 813823. https://doi.org/10.1038/ngeo1955CrossRefGoogle Scholar
Landauer, M., Juhola, S., & Söderholm, M. (2015). Inter-relationships between adaptation and mitigation: A systematic literature review. Climatic Change, 131, 505517. https://doi.org/10.1007/s10584-015-1395-1CrossRefGoogle Scholar
Lausch, A., Borg, E., Bumberger, J., Dietrich, P., Heurich, M., Huth, A., Jung, A., Klenke, R., Knapp, S., Mollenhauer, H., Paasche, H., Paulheim, H., Pause, M., Schweitzer, C., Schmulius, C., Settele, J., Skidmore, A. K., Wegmann, M., Zacharias, S., … Schaepman, M. E. (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, 1120. https://doi.org/10.3390/rs10071120CrossRefGoogle Scholar
Lehmann, C. E. R., Anderson, T. M., Sankaran, M., Higgins, S. I., Archibald, S., Hoffmann, W. A., Hanan, N. P., Williams, R. J., Fensham, R. J., Felfili, J., Hutley, L. B., Ratnam, J., San Jose, J., Montes, R., Franklin, D., Russell-Smith, J., Ryan, C. M., Durigan, G., Hiernaux, P., … Bond, W. J. (2014). Savanna vegetation–fire–climate relationships differ among continents. Science (New York, N.Y.), 343, 548552. https://doi.org/10.1126/science.1247355CrossRefGoogle ScholarPubMed
Lewis, S. L., Mitchard, E. T. A., Prentice, C., Maslin, M., & Poulter, B. (2019). Comment on “The global tree restoration potential”. Science (New York, N.Y.), 366, eaaz0388. https://doi.org/10.1126/science.aaz0388CrossRefGoogle Scholar
Li, F., Lawrence, D. M., & Bond-Lamberty, B. (2017). Impact of fire on global land surface air temperature and energy budget for the 20th century due to changes within ecosystems. Environmental Research Letters, 12, 044014. https://doi.org/10.1088/1748-9326/aa6685CrossRefGoogle Scholar
Liu, Y., Li, Y., Huang, J., Zhu, Q., & Wang, S. (2019). Attribution of the Tibetan Plateau to Northern drought. National Science Review. https://doi.org/10.1093/nsr/nwz191Google ScholarPubMed
Locatelli, B., Pavageau, C., Pramova, E., & Di Gregorio, M. (2015). Integrating climate change mitigation and adaptation in agriculture and forestry: Opportunities and trade-offs. WIREs Climate Change, 6, 585598. https://doi.org/10.1002/wcc.357CrossRefGoogle Scholar
Lu, J. G. (2020). Air pollution: A systematic review of its psychological, economic, and social effects. Current Opinion in Psychology, 32, 5265. https://doi.org/10.1016/j.copsyc.2019.06.024CrossRefGoogle ScholarPubMed
Mahecha, M. D., Gans, F., Brandt, G., Christiansen, R., Cornell, S. E., Fomferra, N., Kraemer, G., Peters, J., Bodesheim, P., Camps-Valls, G., Donges, J. F., Dorigo, W., Estupinan-Suarez, L. M., Gutierrez-Velez, V. H., Gutwin, M., Jung, M., Londoño, M. C., Miralles, D. G., Papastefanou, P., … Reichstein, M. (2020). Earth system data cubes unravel global multivariate dynamics. Earth System Dynamics, 11, 201234. https://doi.org/10.5194/esd-11-201-2020CrossRefGoogle Scholar
Melton, J. R., Wania, R., Hodson, E. L., Poulter, B., Ringeval, B., Spahni, R., Bohn, T., Avis, C. A., Beerling, D. J., Chen, G., Eliseev, A. V., Denisov, S. N., Hopcroft, P. O., Lettenmaier, D. P., Riley, W. J., Singarayer, J. S., Subin, Z. M., Tian, H., Zürcher, S., … Kaplan, J. O. (2013). Present state of global wetland extent and wetland methane modelling: Conclusions from a model inter-comparison project (WETCHIMP). Biogeosciences, 10(2), 753788. https://doi.org/10.5194/bg-10-753-2013CrossRefGoogle Scholar
Meyer, J. L., Strayer, D. L., Wallace, J. B., Eggert, S. L., Helfman, G. S., & Leonard, N. E. (2007). The contribution of headwater streams to biodiversity in river networks. Journal of the American Water Resources Association, 43, 86103. https://doi.org/10.1111/j.1752-1688.2007.00008.xCrossRefGoogle Scholar
Mills, G., Pleijel, H., Malley, C. S., Sinha, B., Cooper, O. R., Schultz, M. G., Neufeld, H. S., Simpson, 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 Athropocene, 6, 4792. http://doi.org/10.1525/elementa.302Google Scholar
Mishra, A. K., Sinha, B., Kumar, R., Barth, M., Hakkim, H., Kumar, V., Kumar, A., Datta, S., Guenther, A., & Sinha, V. (2021). Cropland trees need to be included for accurate model simulations of land-atmosphere heat fluxes, temperature, boundary layer height, and ozone. Science of the Total Environment, 751, 141728. https://doi.org/10.1016/j.scitotenv.2020.141728CrossRefGoogle ScholarPubMed
OECD. (2016). The economic consequences of outdoor air pollution, OECD Publishing, Paris. https://www.oecd.org/environment/indicators-modelling-outlooks/Policy-Highlights-Economic-consequences-of-outdoor-air-pollution-web.pdf (Accessed January 2024).CrossRefGoogle Scholar
Ogle, S. M., Domke, G., Kurz, W. A., Rocha, M. T., Huffman, T., Swan, A., Smith, J. E., Woodall, C., & Krug, T. (2018). Delineating managed land for reporting national greenhouse gas emissions and removals to the United Nations Framework Convention on Climate Change. Carbon Balance and Management, 13, 9. https://doi.org/10.1186/s13021-018-0095-3CrossRefGoogle Scholar
Oh, Y., Zhuang, Q., Welp, L. R., Liu, L., Lan, X., Basu, S., Dlugokencky, E. J., Bruhwiler, L., Miller, J. B., Michel, S. E., Schwietzke, S., Tans, P., Ciais, P., & Chanton, J. P. (2022). Improved global wetland carbon isotopic signatures support post-2006 microbial methane emission increase. Communications Earth & Environment, 3, 159. https://doi.org/10.1038/s43247-022-00488-5CrossRefGoogle Scholar
Pangala, S. R., Enrich-Prast, A., Basso, L. S., Peixoto, R. B., Bastviken, D., Hornibrook, E. R. C., Gatti, L. V., Marotta, H., Calazans, L. S. B., Sakuragui, C. M., Bastos, W. R., Malm, O., Gloor, E., Miller, J. B., & Gauci, V. (2017). Large emissions from floodplain trees close the Amazon methane budget. Nature, 552, 230234. https://doi.org/10.1038/nature24639CrossRefGoogle ScholarPubMed
Paquette, A., & Hargreaves, A. L. (2021). Biotic interactions are more often important at species’ warm versus cool range edges. Ecology Letters, 24, 24272438. https://doi.org/10.1111/ele.13864CrossRefGoogle ScholarPubMed
Pastorello, G., Trotta, C., Canfora, E., Chu, H., Christianson, D., Cheah, Y.-W., Poindexter, C., Chen, J., Elbashandy, A., Humphrey, M., Isaac, P., Polidori, D., Reichstein, M., Ribeca, A., van Ingen, C., Vuichard, N., Zhang, L., Amiro, B., Ammann, C., … Papale, D. (2020). The FLUXNET2015 dataset and the ONEFlux processing pipeline for eddy covariance data. Scientific Data, 7, 225. https://doi.org/10.1038/s41597-020-0534-3CrossRefGoogle ScholarPubMed
Plumptre, A. J., Baisero, D., Belote, R. T., Vázquez-Domínguez, E., Faurby, S., Jȩdrzejewski, W., Kiara, H., Kühl, H., Benítez-López, A., Luna-Aranguré, C., Voigt, M., Wich, S., Wint, W., Gallego-Zamorano, J., & Boyd, C. (2021). Where might we find ecologically intact communities? Frontiers in Forests and Global Change, 4, 626635. https://doi.org/10.3389/ffgc.2021.626635CrossRefGoogle Scholar
Pörtner, H. O., Scholes, R. J., Agard, J., Archer, E., Arneth, A., Bai, X., Barnes, D., Burrows, M., Chan, L., Cheung, W. L., Diamond, S., Donatti, C., Duarte, C., Eisenhauer, N., Foden, W., Gasalla, M. A., Handa, C., Hickler, T., Hoegh-Guldberg, O., … Ngo, H. T. (2021). Scientific outcome of the IPBES-IPCC co-sponsored workshop on biodiversity and climate change; IPBES secretariat, Bonn, Germany. https://doi.org/10.5281/zenodo.4659158CrossRefGoogle Scholar
Poulter, B., Frank, D., Ciais, P., Myneni, R. B., Andela, N., Bi, J., Broquet, G., Canadell, J. G., Chevallier, F., Liu, Y. Y., Running, S. W., Sitch, S., & van der Werf, G. R. (2014). Contribution of semi-arid ecosystems to interannual variability of the global carbon cycle. Nature, 509, 600603. https://doi.org/10.1038/nature13376CrossRefGoogle ScholarPubMed
Retallack, G. J. (2013). Global cooling by grassland soils of the geological past and near future. Annual Review of Earth and Planetary Sciences, 41, 6986. https://doi.org/10.1146/annurev-earth-050212-124001CrossRefGoogle Scholar
Ryan, C. M., Williams, M., & Grace, J. (2011). Above- and belowground carbon stocks in a Miombo woodland landscape of Mozambique. Biotropica, 43, 423432. https://doi.org/10.1111/j.1744-7429.2010.00713.xCrossRefGoogle Scholar
Sankaran, M., Hanan, N. P., Scholes, R. J., Ratnam, J., Augustine, D. J., Cade, B. S., Gignoux, J., Higgins, S. I., Le Roux, X., Ludwig, F., Ardo, J., Banyikwa, F., Bronn, A., Bucini, G., Caylor, K. K., Coughenour, M. B., Diouf, A., Ekaya, W., Feral, C. J., … Zambatis, N. (2005). Determinants of woody cover in African savannas. Nature, 438, 846849. https://doi.org/10.1038/nature04070CrossRefGoogle ScholarPubMed
Saunois, M., Stavert, A. R., Poulter, B., Bousquet, P., Canadell, J. G., Jackson, R. B., Raymond, P. A., Dlugokencky, E. J., Houweling, S., Patra, P. K., Ciais, P., Arora, V. K., Bastviken, D., Bergamaschi, P., Blake, D. R., Brailsford, G., Bruhwiler, L., Carlson, K. M., Carrol, M., … Zhuang, Q. (2020). The Global Methane Budget 2000–2017. Earth System Science Data, 12, 15611623. https://doi.org/10.5194/essd-12-1561-2020CrossRefGoogle Scholar
Schuur, E. A. G., Abbott, B. W., Commane, R., Ernakovich, J., Euskirchen, E., Hugelius, G., Grosse, G., Jones, M., Koven, C., Leshyk, V., Lawrence, D., Loranty, M. M., Mauritz, M., Olefeldt, D., Natali, S., Rodenhizer, H., Salmon, V., Schädel, C., Strauss, J., … Turetsky, M. (2022). Permafrost and climate change: Carbon cycle feedbacks from the warming Arctic. Annual Review of Environment and Resources, 47, 343371. https://doi.org/10.1146/annurev-environ-012220-011847CrossRefGoogle Scholar
Shiklomanov, A. N., Bradley, B. A., Dahlin, K. M., Fox, M., Gough, A., Hoffman, C. M., M Middleton, F. M., Serbin, E., Smallman, S. P., & Smith, L., & K, W. (2019). Enhancing global change experiments through integration of remote-sensing techniques. Frontiers in Ecology and the Environment, 17, 215224. https://doi.org/10.1002/fee.2031CrossRefGoogle Scholar
Suni, T., Guenther, A., Hansson, H. C., Kulmala, M., Andreae, M. O., Arneth, A., Artaxo, P., Blyth, E., Brus, M., Ganzeveld, L., Kabat, P., de Noblet-Ducoudré, N., Reichstein, M., Reissell, A., Rosenfeld, D., & Seneviratne, S. (2015). The significance of land-atmosphere interactions in the Earth system – iLEAPS achievements and perspectives. Anthropocene, 12, 6984. https://doi.org/10.1016/j.ancene.2015.12.001CrossRefGoogle Scholar
Suni, T., Juhola, S., Korhonen-Kurki, K., Käyhkö, J., Soini, K., & Kulmala, M. (2016). National Future Earth platforms as boundary organizations contributing to solutions-oriented global change research. Current Opinion in Environmental Sustainability, 23, 6368. https://doi.org/10.1016/j.cosust.2016.11.011CrossRefGoogle Scholar
Tan, Z. D., Carrasco, L. R., Sutikno, S., & Taylor, D. (2022). Peatland restoration as an affordable nature-based climate solution with fire reduction and conservation co-benefits in Indonesia. Environmental Research Letters, 17, 064028. https://doi.org/10.1088/1748-9326/ac6f6eCrossRefGoogle Scholar
Turetsky, M. R., Kotowska, A., Bubier, J., Dise, N. B., Crill, P., Hornibrook, E. R. C., Minkkinen, K., Moore, T. R., Myers-Smith, I. H., Nykänen, H., Olefeldt, D., Rinne, J., Saarnio, S., Shurpali, N., Tuittila, E.-S., Waddington, J. M., White, J. R., Wickland, K. P., & Wilmking, M. (2014). A synthesis of methane emissions from 71 northern, temperate, and subtropical wetlands. Global Change Biology, 20, 21832197. https://doi.org/10.1111/gcb.12580CrossRefGoogle ScholarPubMed
UNEA. (2022). UNEA resolution 4/14 and 5/2: Sustainable Nitrogen Management, 159th meeting of the Committee of Permanent Representatives to the United Nations Environment Programme. https://wedocs.unep.org/bitstream/handle/20.500.11822/40667/5.a%20UNEA%20resolution%205.2%20-%20Progress%20Sustainable%20Nitrogen%20Management.pdf (Accessed January 2024).Google Scholar
UNEP. (2022). Spreading like Wildfire – The Rising Threat of Extraordinary Landscape Fires. Rapid Response Assessment. United Nations Environment Programme (Edited by Sullivan, A., Baker, E. and Kurvits, T.), Nairobi, Kenya. https://www.unep.org/resources/report/spreading-wildfire-rising-threat-extraordinary-landscape-fires (Accessed January 2024).Google Scholar
UNFCCC. (2015). Adoption of the Paris Agreement, FCCC/CP/2015/L.9/Rev. 1. http://unfccc.int/resource/docs/2015/cop21/eng/l09r01.pdf (Accessed January 2024).Google Scholar
USEPA. (2010). Methane and Nitrous Oxide Emissions From Natural Sources, US Environmental Protection Agency (EPA 430-R-10-001). https://nepis.epa.gov/Exe/ZyPDF.cgi/P100717T.PDF?Dockey=P100717T.PDF (Accessed January 2024).Google Scholar
van der Hel, S. (2016). New science for global sustainability? The institutionalisation of knowledge co-production in Future Earth. Environmental Science & Policy, 61, 165175. https://doi.org/10.1016/j.envsci.2016.03.012CrossRefGoogle Scholar
Van Dingenen, R., Dentener, F. J., Raes, F., Krol, M. C., Emberson, L., & Cofala, J. (2009). The global impact of ozone on agricultural crop yields under current and future air quality legislation. Atmospheric Environment, 43, 604618. https://doi.org/10.1016/j.atmosenv.2008.10.033CrossRefGoogle Scholar
Wang, X., Wu, Z., Zhang, Q., Cohen, J., & Pang, J. (2017). Impact of urbanization on regional climate and Air quality in China. In Bouarar, I., Wang, X., & Brasseur, G. (Eds.), Air pollution in Eastern Asia: An integrated perspective. ISSI scientific report series, vol 16. Springer. https://doi.org/10.1007/978-3-319-59489-7_22Google Scholar
Zhang, Z., Poulter, B., Feldman, A. F., Ying, Q., Ciais, P., Peng, S., & Li, X. (2023). Recent intensification of wetland methane feedback. Nature Climate Change, 13, 430433. https://doi.org/10.1038/s41558-023-01629-0CrossRefGoogle Scholar
Figure 0

Figure 1. Key risks to terrestrial and freshwater ecosystems from climate change (Figure 2.11 in IPCC AR6 WGII [IPCC, 2022a]).

Figure 1

Figure 2. Schematic of the three focal systems of interest to iLEAPS, their overlap, the three key cross-cutting themes, and the land–atmosphere processes involved.

Figure 2

Figure 3. Schematic diagram of the impacts of urbanization on climate and air quality (adapted from Wang et al. [2017]).

Supplementary material: File

Hayman et al. supplementary material 1

Hayman et al. supplementary material
Download Hayman et al. supplementary material 1(File)
File 54 KB
Supplementary material: File

Hayman et al. supplementary material 2

Hayman et al. supplementary material
Download Hayman et al. supplementary material 2(File)
File 277.1 KB