Hostname: page-component-848d4c4894-x5gtn Total loading time: 0 Render date: 2024-06-02T05:57:34.476Z Has data issue: false hasContentIssue false

The present, past and future of blue carbon

Published online by Cambridge University Press:  08 June 2023

Kerrylee Rogers*
Affiliation:
School of Earth, Atmospheric and Life Sciences, University of Wollongong, Wollongong, NSW 2522, Australia
Jeffrey J. Kelleway
Affiliation:
School of Earth, Atmospheric and Life Sciences, University of Wollongong, Wollongong, NSW 2522, Australia
Neil Saintilan
Affiliation:
School of Natural Sciences, Macquarie University, North Ryde, NSW 2109, Australia
*
Corresponding author: Kerrylee Rogers; Email: kerrylee@uow.edu.au
Rights & Permissions [Opens in a new window]

Abstract

Blue carbon is identified as a natural climate solution as it provides multiple ecosystem services, including climate mitigation, adaptation, and other co-benefits. There remain ongoing challenges for blue carbon as a natural climate solution, particularly as blue carbon ecosystems are at risk from climate change. Concepts of uniformitarianism were applied to consider how the present and past behaviour of blue carbon ecosystems can inform decision-makers of blue carbon risks. Climate change may increase the capture and storage of blue carbon in the short to medium term; this is largely due to negative feedbacks between elevated atmospheric carbon dioxide and temperature, and supplemented by natural processes of sediment supply and accumulation. Opportunities for retreat and increasing carbon storage as sea levels rise are likely to be greater where sea level has a longer history of relative stability, largely in the Southern Hemisphere. Landward retreat will be crucial where millennia of sea-level rise have limited the capacity for in situ blue carbon additionality; this may be thwarted by highly developed coastal zones and coastal squeeze effects. Negative feedbacks may fail under higher emissions, greater warming and rates of sea-level rise exceeding ~5–7 mm yr.−1; this tipping point may be surpassed within the next century under a high emissions scenario. Retreat of blue carbon ecosystems to higher elevations where they are afforded protection from the effects of sea-level rise will be critical for blue carbon additionality. Carbon markets are prepared to incentivise restoration of blue carbon ecosystems as they adapt to climate change; however, knowledge gaps remain, particularly regarding the behaviour of blue carbon ecosystems in the global south. Given the momentum in blue carbon research, scientists and practitioners are well placed to continue addressing blue carbon risks.

Type
Review
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
© The Author(s), 2023. Published by Cambridge University Press

Impact statement

The role of coastal ecosystems in sequestering atmospheric carbon has been demonstrated and efforts are underway to leverage this service for climate change mitigation. Focussing largely on mangroves and saltmarshes, this paper applies concepts of uniformitarianism to consider how the present behaviour of blue carbon ecosystems (BCEs), and their response to varying sea level and climate change in the past, provides knowledge of BCE futures. There is evidence BCEs are being modified by climate change, and losses are compounded by ongoing clearance and degradation. Consequently, atmospheric warming, sea-level rise and human modifications have important implications for greenhouse gas fluxes from BCEs. Observations of the response of BCEs to sea-level rise indicate some capacity to adapt by accumulating mineral and organic material and increasing substrate elevations. Contributions of organic matter to substrates as sea level rises, coupled with landward retreat of BCEs to maintain their intertidal position, bolsters capacity to sequester atmospheric carbon within substrates and biomass. However, evidence of the response of mangroves and saltmarshes to sea-level rise over the Holocene indicate a threshold for adaptation that is exceeded when sea-level rises at ~5–7 mm yr.−1. This rate of sea-level rise is anticipated to be surpassed by the end of the 21st century unless deep cuts to atmospheric carbon concentrations occur rapidly. Accordingly, the fate of BCEs remains uncertain. However, it is probable mangrove forests and saltmarshes high in the tidal frame will transition towards lower tidal positions as sea level rises and may survive for some time, offering some confidence when trading carbon with a 25-year permanence timeframe. Landward retreat will be critical for blue carbon additionality, and this will require concerted management to minimise coastal squeeze and preserve BCE services. This uncertainty should be accommodated when considering the permanence of blue carbon in financial markets.

Introduction

Blue carbon is a collective term referring to the carbon associated with marine and coastal ecosystems, and includes all fluxes and stores that are biologically driven (Bindoff et al., Reference Bindoff, Cheung, Kairo, Arístegui, Guinder, Hallberg, Pörtner, Roberts and Masson-Delmotte2019). Similar to carbon sequestered in terrestrial forests, blue carbon has piqued the interest of practitioners seeking to mitigate climate change by enhancing carbon storage within natural ecosystems and improving the provision of ecosystem services (Macreadie et al., Reference Macreadie, Costa, Atwood, Friess, Kelleway, Kennedy, Lovelock, Serrano and Duarte2021). This interest is based on the high carbon storage potential of many blue carbon ecosystems (BCEs), a potential that is reported to be much higher on a unit area basis than other ecosystem-based climate solutions (Donato et al., Reference Donato, Kauffman, Murdiyarso, Kurnianto, Stidham and Kanninen2011; Pendleton et al., Reference Pendleton, Donato, Murray, Crooks, Jenkins, Sifleet, Craft, Fourqurean, Kauffman and Marbà2012; Duarte et al., Reference Duarte, Losada, Hendriks, Mazarrasa and Marba2013). BCEs are typically vegetated with mangroves, saltmarshes (also termed tidal marshes) and seagrasses, and to a lesser extent macroalgae, cyanobacteria and supratidal forests (Duarte et al., Reference Duarte, Losada, Hendriks, Mazarrasa and Marba2013; Raven, Reference Raven2018; Bindoff et al., Reference Bindoff, Cheung, Kairo, Arístegui, Guinder, Hallberg, Pörtner, Roberts and Masson-Delmotte2019; Lovelock and Duarte, Reference Lovelock and Duarte2019) (Figure 1). Carbon is drawn from the atmosphere via photosynthesis and stored within living biomass at a concentration of 40–50% of the mass, a value that is reasonably consistent among plants (Ma et al., Reference Ma, He, Tian, Zou, Yan, Yang, Zhou, Huang, Shen and Fang2018). Blue carbon is partitioned into above- and below-ground biomass and the soil carbon pool. Above-ground biomass is typically estimated from allometric equations, initially derived from destructive measurements that relate vegetation structure to mass (Thursby et al., Reference Thursby, Chintala, Stetson, Wigand and Champlin2002; Komiyama et al., Reference Komiyama, Ong and Poungparn2008; Radabaugh et al., Reference Radabaugh, Powell, Bociu, Clark and Moyer2017), or by applying remote sensing techniques to extrapolate spatial relationships (Pham et al., Reference Pham, Xia, Ha, Bui, Le and Tekeuchi2019; Sani et al., Reference Sani, Hashim and Hossain2019). The below-ground component is somewhat more difficult to quantify as substrates contain both living biomass and dead organic material that has accumulated over decades to thousands of years, as evident from radiocarbon dating of BCEs (Horton et al., Reference Horton, Shennan, Bradley, Cahill, Kirwan, Kopp and Shaw2018; Saintilan et al., Reference Saintilan, Khan, Ashe, Kelleway, Rogers, Woodroffe and Horton2020; Sefton et al., Reference Sefton, Woodroffe, Ascough, Friess and Sidik2021).

Figure 1. Global mapped distribution and existing estimates of carbon cycling parameters of (A) saltmarsh and (B) mangrove; (C) modelled distribution of seagrass; and (D) genus richness of benthic marine macroalgae. All values are global mean values ±1 standard deviation (where available) unless otherwise specified. Belowground carbon stocks are estimated to 1 m depth. CAR = (surface) carbon accumulation rate; NPP = net ecosystem primary productivity; SE = 1 standard error. Note that for macroalgae, CAR is replaced by estimates of carbon burial in situ (i.e. in algal beds) and exported particulate organic carbon buried in shelf sediments. Map data sources: saltmarsh (Mcowen et al., Reference Mcowen, Weatherdon, Bochove, Sullivan, Blyth, Zockler, Stanwell-Smith, Kingston, Martin, Spalding and Fletcher2017); mangrove (Bunting et al., Reference Bunting, Rosenqvist, Lucas, Rebelo, Hilarides, Thomas, Hardy, Itoh, Shimada and Finlayson2018); seagrass (Jayathilake and Costello, Reference Jayathilake and Costello2018); macroalgae (Kerswell, Reference Kerswell2006). Carbon data sources: a (Rogers et al., Reference Rogers, Kelleway, Saintilan, Megonigal, Adams, Holmquist, Lu, Schile-Beers, Zawadzki, Mazumder and Woodroffe2019a); b (Pendleton et al. Reference Pendleton, Donato, Murray, Crooks, Jenkins, Sifleet, Craft, Fourqurean, Kauffman and Marbà2012a); c (Duarte and Cebrian, Reference Duarte and Cebrian1996); d (Wang et al., Reference Wang, Sanders, Santos, Tang, Schuerch, Kirwan, Kopp, Zhu, Li and Yuan2021); e (Ouyang and Lee, Reference Ouyang and Lee2014); f (Atwood et al., Reference Atwood, Connolly, Almahasheer, Carnell, Duarte, Ewers Lewis, Irigoien, Kelleway, Lavery, Macreadie, Serrano, Sanders, Santos, Steven and Lovelock2017); g; h (Alongi, Reference Alongi2012); i; j (McLeod et al., Reference McLeod, Chmura, Bouillon, Salm, Björk, Duarte, Lovelock, Schlesinger and Silliman2011); k (Krause-Jensen and Duarte, Reference Krause-Jensen and Duarte2016).

The enhanced capacity for storage is dependent upon rates of carbon addition exceeding loss of carbon via the decomposition of organic material, and there is increasing agreement that this should also exceed in situ carbonate production (Saderne et al., Reference Saderne, Geraldi, Macreadie, Maher, Middelburg, Serrano, Almahasheer, Arias-Ortiz, Cusack and BDJNc2019). High net primary production from in situ vegetation underpins the supply of organic matter to substrates, mostly from root material (Saintilan et al., Reference Saintilan, Rogers, Mazumder and Woodroffe2013; Xiong et al., Reference Xiong, Liao and Wang2018). However, organic matter transported on tides can also become trapped and sequestered into substrates, and there is an increasing need to discriminate the varying role of autochthonous and allochthonous sources (Saintilan et al., Reference Saintilan, Rogers, Mazumder and Woodroffe2013; Canuel and Hardison, Reference Canuel and Hardison2016; Van de Broek et al., Reference Van de Broek, Vandendriessche, Poppelmonde, Merckx, Temmerman and Govers2018; Macreadie et al., Reference Macreadie, Anton, Raven, Beaumont, Connolly, Friess, Kelleway, Kennedy, Kuwae, Lavery, Lovelock, Smale, Apostolaki, Atwood, Baldock, Bianchi, Chmura, Eyre, Fourqurean, Hall-Spencer, Huxham, Hendriks, Krause-Jensen, Laffoley, Luisetti, Marbà, Masque, KJ, Megonigal, Murdiyarso, Russell, Santos, Serrano, Silliman, Watanabe and Duarte2019). Periodic inundation by saline tidal waters creates anaerobic conditions in saturated substrates that supress microbial activity and slow the decomposition of sequestered organic material (Duarte et al., Reference Duarte, Losada, Hendriks, Mazarrasa and Marba2013; Spivak et al., Reference Spivak, Sanderman, Bowen, Canuel and Hopkinson2019). Addition of mineral sediments supplied by tides serves to trap sequestered organic material (Spivak et al., Reference Spivak, Sanderman, Bowen, Canuel and Hopkinson2019) and saline substrates hamper methanogenic processes (Poffenbarger et al., Reference Poffenbarger, Needelman and Megonigal2011). While greenhouse gas emissions are not fully supressed (Rosentreter et al., Reference Rosentreter, Maher, Erler, Murray and Eyre2018), the general outcome is physicochemical conditions that favour slow decomposition of organic material, long-term preservation of a portion of fixed carbon within substrates, and limited release of powerful greenhouse gases (e.g. methane and nitrous oxide) to the atmosphere (McKee et al., Reference McKee, Cahoon and Feller2007; McLeod et al., Reference McLeod, Chmura, Bouillon, Salm, Björk, Duarte, Lovelock, Schlesinger and Silliman2011; Kroeger et al., Reference Kroeger, Crooks, Moseman-Valtierra and Tang2017). Storage may be further enhanced when coastal processes operate to ensure that space within substrates for carbon storage continues to be available, and this appears to be strongly influenced by rates of sediment supply, sedimentation, and coastal evolution in the context of changing sea levels (Rogers et al., Reference Rogers, Kelleway, Saintilan, Megonigal, Adams, Holmquist, Lu, Schile-Beers, Zawadzki, Mazumder and Woodroffe2019a). Together these conditions mean that BCEs can store orders of magnitude more carbon within their substrates than other terrestrial ecosystems, estimated to be in the order of 0.4–6.5 Pg. C in the upper 1 m of saltmarsh substrates globally, 9.4–10.4 Pg. C for mangrove forest substrates and 4.2–8.4 Pg. C for seagrass substrates (Duarte et al., Reference Duarte, Losada, Hendriks, Mazarrasa and Marba2013).

Estimates of carbon storage in BCEs are typically determined based on their current distribution. At the coarsest level, a central measure (e.g. mean, median) of carbon concentration is multiplied by BCE extent to estimate carbon storage in various components (above- and below-ground biomass, soil carbon pool) (Howard et al., Reference Howard, Hoyt, Isensee, Pidgeon and Telszewski2014). As can be seen in Figure 1 such projections to a global scale will be subject to large uncertainties associated with the high variability, across multiple spatial scales, in estimates within each ecosystem type; spatial biases in data availability also influence confidence in global projections.

In spite of this increasing recognition of spatial and temporal variation in carbon storage, it is probable that additional soil organic carbon originating from BCEs is preserved within coastal floodplains and on continental shelfs where conditions are now no longer favourable for BCEs, but where long-term preservation may have occurred over millennia as coastal landscapes evolved (Hanebuth et al., Reference Hanebuth, Stattegger and Grootes2000; Grindrod, Reference Grindrod, Kershaw, David, Tapper, Penny and Brown2002; Rogers et al., Reference Rogers, Macreadie, Kelleway and Saintilan2019b). Additionally, there is also an imprint of direct human impacts on BCEs, with losses largely due to land cover change and gains largely due to restoration activities leading to a net decline in tidal wetland extent (inclusive of tidal flats, mangrove forests and tidal marshes) of ~4,000 km2 between 1999 and 2019 (Murray et al., Reference Murray, Worthington, Bunting, Duce, Hagger, Lovelock, Lucas, Saunders, Sheaves and Spalding2022). There is evidence that this rate of loss is diminishing (Friess et al., Reference Friess, Yando, Abuchahla, Adams, Cannicci, Canty, Cavanaugh, Connolly, Cormier and Dahdouh-Guebas2020; Campbell et al., Reference Campbell, Fatoyinbo, Goldberg and Lagomasino2022) and substantial gains to tidal wetland extent, in the order of 9,700 km2, have been related to the success of restoration activities and natural expansion (Friess et al., Reference Friess, Rogers, Lovelock, Krauss, Hamilton, Lee, Lucas, Primavera, Rajkaran and Shi2019; Murray et al., Reference Murray, Worthington, Bunting, Duce, Hagger, Lovelock, Lucas, Saunders, Sheaves and Spalding2022). Accordingly, the current distribution of BCEs indicates where current storage and additionality occurs but does not indicate storage that occurred prior to changes in land cover at millennia timescales or over the record of Earth observations. This has important implications for the future of blue carbon, and it is probable that the geographic distribution of BCEs will continue to change as sea level rises, coasts evolve, and humans living in the coastal zone adapt to a new configuration of the coast and BCE distribution.

To characterise the future of blue carbon as a natural climate solution, we apply concepts of uniformitarianism to consider how the contemporary distribution and behaviour of blue carbon can inform our understanding of the long-term evolution of coastal carbon storage. In this regard, we consider processes that influence blue carbon over the observational record (i.e. the present) are the same as those that have operated for millennia (i.e. the past), and, by extrapolation, consider blue carbon futures. We specifically consider the evidence preserved in stratigraphic records of the evolution of BCEs and present the case that blue carbon storage is related to coastal evolution and influenced by sediment supply and Holocene sea-level change. We also recognise that coastal evolution and the influence of humans on coastal landscapes means that the response of BCEs in the past is not a direct analogue for their future adaptation to anticipated sea-level rise and climate change (Woodroffe and Murray-Wallace, Reference Woodroffe and Murray-Wallace2012). Nevertheless, present-day coastal landscapes are an archive of information that can be used to parameterise models projecting the response of BCEs to environmental change; thereby providing the critical information needed to inform coastal zone planning and decision-making, improve the resilience of BCEs and ensure their long-term application as a nature-based solution that contributes to climate mitigation efforts.

The PRESENT: Blue carbon in coastal landscapes

Mangroves and saltmarsh typically occupy the upper half of the intertidal zone, with mangrove forests dominating intertidal shorelines of the tropics and saltmarshes dominating intertidal shorelines of temperate zones (Figure 2). There is considerable overlap in the latitudinal distribution of mangroves and saltmarshes (Figure 1), with mangroves generally limited to ocean temperatures that exceed 20 °C during the coldest month (West, Reference West1956; Quisthoudt et al., Reference Quisthoudt, Schmitz, Randin, Dahdouh-Guebas, Robert and Koedam2012; Osland et al., Reference Osland, Enwright, Day, Gabler, Stagg and Grace2016), while saltmarsh distribution is influenced by substrate salinity (Bertness et al., Reference Bertness, Gough and Shumway1992; Silvestri et al., Reference Silvestri, Defina and Marani2005). Where salinity is very high and salts concentrate hypersaline flats/sandflats or sabkha predominate, often with cyanobacterial mats, and saltmarsh vegetation is sparse; in the tropics where rainfall is high, mangroves predominate in the upper half of the intertidal zone (Rogers and Woodroffe, Reference Rogers, Woodroffe, Masselink and Gehrels2014). Most seagrass species occupy fully inundated substrates (i.e. subtidal elevations) where water clarity is a significant control on productivity (Madsen et al., Reference Madsen, Chambers, James, Koch and Westlake2001). Where hydrodynamic conditions allow, a diversity of seagrass genera may also occupy the lower intertidal niche (i.e. below MSL) (Björk et al., Reference Björk, Uku, Weil and Beer1999). Depth-constrained species of seagrass and macroalgae can accumulate organic material within substrates as they adjust to changing water levels (e.g. Zostera spp., Halophila spp. and Phyllospadix spp.) (Koch, Reference Koch2001; Madsen et al., Reference Madsen, Chambers, James, Koch and Westlake2001). While seagrass meadows may be an exceptional carbon source for sequestration elsewhere, their capacity for in situ blue carbon storage is largely limited to that stored within the living biomass and some detritus. Increased sequestration is therefore largely dependent upon an increase in the lateral extent of seagrass meadows (Greiner et al., Reference Greiner, McGlathery, Gunnell and McKee2013) and should be balanced against the additional CO2 that is released by carbonate sediment production (Howard et al., Reference Howard, Creed, Aguiar and Fourqurean2018). While net ecosystem primary production may be high for macroalgae, with potential importance for the export of carbon to other environments, in situ burial of carbon is limited (Figure 1; Krause-Jensen and Duarte, Reference Krause-Jensen and Duarte2016). For these reasons, this review focuses on mangrove and saltmarsh blue carbon futures.

Figure 2. Profiles of BCE landscapes indicating (A) accommodation space, delimited by the highest astronomical tide, basement or bedrock geology, and hydrodynamic conditions favourable for mineral and organic matter accumulation (modified from Rogers (Reference Rogers2021)); and (B) the range of techniques that can be used to observe and measure changes in substrate volume, mineral and organic matter accumulation, and position within the tidal frame, with specific focus on changing tidal position with sea-level rise, as per Allen (Reference Allen2000).

Geomorphological settings occupied by BCEs are delimited to areas where low-energy intertidal substrates support the establishment and maintenance of salt-tolerant vegetation. Globally, deltas are hotspots for BCEs due to high rates of sediment supply promoting the development of broad intertidal environments where resource availability is high (Rovai et al., Reference Rovai, Twilley, Castañeda-Moya, Riul, Cifuentes-Jara, Manrow-Villalobos, Horta, Simonassi, Fonseca and Pagliosa2018; Worthington et al., Reference Worthington, Zu Ermgassen, Friess, Krauss, Lovelock, Thorley, Tingey, Woodroffe, Bunting and Cormier2020; Murray et al., Reference Murray, Worthington, Bunting, Duce, Hagger, Lovelock, Lucas, Saunders, Sheaves and Spalding2022). Along tide-dominated coasts, favourable conditions may arise on the open coast where tidally borne sediments can accumulate, or within the intertidal zone of tide-dominated estuaries. Along wave-dominated coasts, barriers at estuary entrances dampen wave energy and terrigenous sediments supplied from catchments via distributaries contribute to intertidal development of fluvial deltas, and to a lesser extent marine sediment delivered by tides contribute to the development of flood-tide deltas that support BCEs. These fluvial and floodtide deltas associated with wave-dominated estuaries are also hotspots for blue carbon due to the development of low-gradient intertidal substrates (Kelleway et al., Reference Kelleway, Saintilan, Macreadie and Ralph2016a).

Global and continental scale analyses of above-ground biomass of mangroves and saltmarshes variably highlight the role of climatic factors that vary with latitude, proposing that optimal temperature and higher rainfall favours productivity and carbon addition to plants within BCEs (Kirwan and Mudd, Reference Kirwan and Mudd2012; Rovai et al., Reference Rovai, Riul, Twilley, Castañeda‐Moya, Rivera‐Monroy, Williams, Simard, Cifuentes‐Jara, Lewis and Crooks2016; Sanders et al., Reference Sanders, Maher, Tait, Williams, Holloway, Sippo and Santos2016). For example, analyses of mangrove heights and biomass using the global shuttle radar topography mission altimetry dataset highlight the role of precipitation, temperature and cyclone frequency, explaining 74% of global trends in mangrove canopy height (Simard et al., Reference Simard, Fatoyinbo, Smetanka, Rivera-Monroy, Castañeda-Moya, Thomas and Van der Stocken2019). Yet, these are also factors that modify substrate salinity and may promote decomposition of organic material (Chmura et al., Reference Chmura, Anisfeld, Cahoon and Lynch2003; Kirwan et al., Reference Kirwan, Guntenspergen and Langley2014; Mueller et al., Reference Mueller, Schile-Beers, Mozdzer, Chmura, Dinter, Kuzyakov, de Groot, Esselink, Smit and D’Alpaos2018). The outcome may be that once living standing stock reaches a threshold biomass, additions to the living biomass are offset by losses to the standing stock (Chmura et al., Reference Chmura, Anisfeld, Cahoon and Lynch2003).

Observations of mangrove above-ground biomass addition in forestry plots indicate that individual tree growth will asymptote at a threshold height and biomass addition is largely limited to small increments to woody components (i.e. thickening of trunks and stems) (Jin-Eong et al., Reference Jin-Eong, Khoon and Clough1995; Osland et al., Reference Osland, Feher, Spivak, Nestlerode, Almario, Cormier, From, Krauss, Russell and Alvarez2020; Alongi, Reference Alongi2020b). It is this asymptotic nature of above-ground biomass addition over time that has led to many forestry-based carbon offsetting schemes having a minimum commitment period of at least 20 years before harvesting can occur (Galik et al., Reference Galik, Baker, Daigneault and Latta2022) and this has translated into voluntary methods for blue carbon offsetting (Lovelock et al., Reference Lovelock, Adame, Bradley, Dittmann, Hagger, Hickey, Hutley, Jones, Kelleway and Lavery2022a). This aligns with the period over which biomass addition accelerates as plants establish and the rate of carbon sequestration is high. When mangrove forests and saltmarshes have reached their threshold capacity for standing above-ground biomass, then increases in the standing living stock are largely achieved by lateral increases in extent as increases in plant density in mature forests will be resource limited. Increases in below-ground biomass are presumed to be limited by vertical space within substrates for net biomass additionality and addition of biomass will increasingly be offset by decomposition of below-ground biomass as substrates asymptote towards higher elevations; that is unless relative sea-level rise creates more vertical space for below-ground storage. Critically, lateral increases in extent are constrained by the availability of land where conditions are favourable within the intertidal and supratidal zone. In addition, although intense cyclones and storms are reported to have a return interval of approximately 20 years (Elsner et al., Reference Elsner, Jagger and Tsonis2006), aligning with the commitment period for restoration projects before harvesting can occur (Galik et al., Reference Galik, Baker, Daigneault and Latta2022), the feasibility of restoration projects in regions with a propensity for cyclone activity may decrease should projected increases in the frequency and intensity of major storms eventuate (IPCC, Reference Masson-Delmotte, Zhai and Pirani2021).

Partitioning of mangrove biomass between above-ground and below-ground differs from terrestrial forests in that a relatively high proportion is allocated to below-ground root systems (Saintilan, Reference Saintilan1997; Lichacz et al., Reference Lichacz, Hardiman and Buckney2009), a possible adaptation to saline conditions (Ball, Reference Ball1988). However, below-ground biomass is somewhat difficult to determine because decomposition is limited by saline and anaerobic conditions of tidally inundated substrates, and differentiating living and dead components of below-ground biomass is difficult (Adame et al., Reference Adame, Cherian, Reef and Stewart-Koster2017). The proportion of mass allocated to above and below-ground components is influenced by environmental conditions, most notable is variation arising from soil water salinity, observed in both laboratory (Ball, Reference Ball1988; Ball, Reference Ball2002) and field studies (Saintilan, Reference Saintilan1997), but also soil water nutrient conditions (Darby and Turner, Reference Darby and Turner2008) and atmospheric CO2 concentrations (Langley et al., Reference Langley, McKee, Cahoon, Cherry and Megonigal2009). Observations of below-ground root addition have been undertaken using root ingrowth bags and ‘marsh organ’ experiments, indicating rapid root development (Muhammad-Nor et al., Reference Muhammad-Nor, Huxham, Salmon, Duddy, Mazars-Simon, Mencuccini, Meir and Jackson2019; Kihara et al., Reference Kihara, Dannoura and Ohashi2022), an observation supported by repeat measures of root biomass (Lamont et al., Reference Lamont, Saintilan, Kelleway, Mazumder and Zawadzki2020).

While the addition of carbon to living biomass makes an important contribution to carbon drawdown from the atmosphere, the substrates of BCEs are typically the largest carbon pool within BCEs with soil organic carbon to a depth of 1 m estimated to comprise 77% of the total global mangrove stock and 95% of the total global saltmarsh stock (Alongi, Reference Alongi2020a). Constraining controls on global below-ground carbon storage has been more elusive, and variably related to edaphic conditions associated with climate and substrate salinity (Chmura et al., Reference Chmura, Anisfeld, Cahoon and Lynch2003; Kirwan and Mudd, Reference Kirwan and Mudd2012; Sanders et al., Reference Sanders, Maher, Tait, Williams, Holloway, Sippo and Santos2016; Rovai et al., Reference Rovai, Twilley, Castañeda-Moya, Riul, Cifuentes-Jara, Manrow-Villalobos, Horta, Simonassi, Fonseca and Pagliosa2018; Sanderman et al., Reference Sanderman, Hengl, Fiske, Solvik, Adame, Benson, Bukoski, Carnell, Cifuentes-Jara and Donato2018). It took some time for the role of sea-level rise to be linked to soil organic carbon accumulation (Wang et al., Reference Wang, Lu, Sanders and Tang2019; Rogers et al., Reference Rogers, Kelleway, Saintilan, Megonigal, Adams, Holmquist, Lu, Schile-Beers, Zawadzki, Mazumder and Woodroffe2019a); this is surprising given the geographic position of intertidal BCEs near mean sea level and the well-established influence of sea-level rise on coastal geomorphology. Increasing mangrove extent with relative sea-level rise and warmer temperatures over the past few decades has been observed and is particularly notable in Australia, Brazil, the Gulf and Atlantic US Coastline, Mexico, and South Africa where mangroves have expanded landward to higher elevations and/or to more poleward positions (Saintilan et al., Reference Saintilan, Wilson, Rogers, Rajkaran and Krauss2014; Godoy and de Lacerda, Reference Godoy and de Lacerda2015; Ximenes et al., Reference Ximenes, Maeda, Arcoverde and Dahdouh-Guebas2016; Osland et al., Reference Osland, Feher, Griffith, Cavanaugh, Enwright, Day, Stagg, Krauss, Howard, Grace and Rogers2017). Increases in soil organic carbon storage have occurred in consort (Kelleway et al., Reference Kelleway, Saintilan, Macreadie, Skilbeck, Zawadzki and Ralph2016b; Simpson et al., Reference Simpson, Stein, Osborne and Feller2019), implying that links between soil organic carbon storage and sea level are partly mediated by vegetation change.

Increases in organic carbon accumulation have been measured in wetlands subject to increased rates of sea-level rise in southwest Florida (Breithaupt et al., Reference Breithaupt, Smoak, Bianchi, Vaughn, Sanders, Radabaugh, Osland, Feher, Lynch and Cahoon2020), and eastern Australia (Marx et al., Reference Marx, Knight, Dwyer, Child, Hotchkis and Zawadzki2020). Extreme rapid subsidence beneath a coastal wetland in Australia, in the order of 1 m, has served as a natural laboratory for observing the influence of relative sea-level rise on carbon storage and addition to mangrove and saltmarsh ecosystems (Rogers et al., Reference Rogers, Kelleway, Saintilan, Megonigal, Adams, Holmquist, Lu, Schile-Beers, Zawadzki, Mazumder and Woodroffe2019a). Here soil organic carbon addition accelerated following an increase in relative sea level. Following subsidence, conditions in the former higher elevation saltmarsh became favourable for mangroves, which rapidly established a deeper root network, supplementing the soil organic carbon pool. The submerged mangrove forest was inundated more frequently, providing more opportunities for carbon-rich tidally borne sediments to accumulate and increase the rate of carbon addition.

The great mangrove forests and saltmarsh plains prior to widespread human-driven change to the coastal zone have been reduced to remnants following decades of clearance for aquaculture, agriculture, coastal developments, and tidal obstructions (Gedan et al., Reference Gedan, Silliman and Bertness2009; Friess et al., Reference Friess, Rogers, Lovelock, Krauss, Hamilton, Lee, Lucas, Primavera, Rajkaran and Shi2019; Goldberg et al., Reference Goldberg, Lagomasino, Thomas and Fatoyinbo2020). This is compounded by conversion to open water arising from land subsidence following groundwater and hydrocarbon extraction and/or associated with diminishing sediment supply arising from damming and diversions. Losses in recent decades of mangrove forests (0.7–3.0% yr.−1), saltmarshes (1.0–2.0% yr.−1) and seagrass meadows (0.4–2.6% yr.−1) have contributed an estimated 0.15–1.02 Pg of CO2 emission per year (Pendleton et al., Reference Pendleton, Donato, Murray, Crooks, Jenkins, Sifleet, Craft, Fourqurean, Kauffman and Marbà2012) (by comparison, 3–19% of emissions from deforestation). These losses reverse decades of carbon sequestration within biomass and millennia of sequestration from substrates. In particular methane flux, a greenhouse gas with ~28 times higher global warming potential than carbon dioxide (Foster et al., Reference Forster, Storelvmo, Armour, Collins, Dufresne, Frame, Lunt, Mauritsen, Palmer, Watanabe, Wild, Zhang, Masson-Delmotte, Zhai, Pirani, Connors, Péan, Berger, Caud, Chen, Goldfarb, Gomis, Huang, Leitzell, Lonnoy, Matthews, Maycock, Waterfield, Yelekçi, Yu and Zhou2021), is markedly higher following substrate disturbance. A global review of methane emissions arising from conversion of mangroves, saltmarshes, seagrasses and tidal flats for coastal aquaculture estimated a rise in methane emissions per area 7–430 times higher than emissions from non-converted coastal habitats (Rosentreter et al., Reference Rosentreter, Borges, Deemer, Holgerson, Liu, Song, Melack, Raymond, Duarte and Allen2021b). Fortunately, there is evidence that this trend of declining BCE extent is slowing (Pendleton et al., Reference Pendleton, Donato, Murray, Crooks, Jenkins, Sifleet, Craft, Fourqurean, Kauffman and Marbà2012; Friess et al., Reference Friess, Rogers, Lovelock, Krauss, Hamilton, Lee, Lucas, Primavera, Rajkaran and Shi2019; Friess et al., Reference Friess, Yando, Abuchahla, Adams, Cannicci, Canty, Cavanaugh, Connolly, Cormier and Dahdouh-Guebas2020), and losses are being offset by the creation of new wetlands (Murray et al., Reference Murray, Worthington, Bunting, Duce, Hagger, Lovelock, Lucas, Saunders, Sheaves and Spalding2022).

Conceptualising blue carbon accommodation space

Accommodation is a term used to define the three-dimensional space available for mineral sediments and soil organic matter (Jervey, Reference Jervey, Wilgus, Hastings and Posamentier1988) (Figure 2A). The maximum elevation of tidal inundation delimits both the landward extent that tidally borne mineral and organic matter can accumulate, and delimits the zone supporting living mangrove and saltmarsh vegetation and in situ soil organic matter contributions (Rogers, Reference Rogers2021). In situ contributions are also delimited at the seaward margin and along tidal creeks by the low-energy hydrodynamic conditions required for vegetation establishment and ongoing survival; in these locations, it is only detrital material that can accumulate within sediments. Initially, bedrock or basement geology delimits the zone in which sediments can accumulate, but as accommodation becomes increasingly infilled, or ‘realised’, via the accumulation of mineral and organic material, substrate elevations increase and become progressively terrestrialised and exposed to the oxidising conditions underpinning aerobic processes of soil organic matter decomposition. In these circumstances, an increase in available accommodation, either via autocompaction of sediments that have accumulated within the ‘realised’ accommodation, subsidence of the basement, or sea-level rise, is required to reinstate tidal inundation, preserve the niche of BCEs, minimise decomposition of organic material by processes of oxidation or methanogenesis and provide new space for additional soil organic matter.

Observations of mangrove and saltmarsh substrate elevation changes using techniques such as surface elevation tables, marker horizons and radiometric dating (Figure 2B) confirm that sedimentation and surface elevation gain are proportional to position in the tidal frame, reflecting the influence of accommodation on accumulation of mineral and organic material in substrates (Webb et al., Reference Webb, Friess, Krauss, Cahoon, Guntenspergen and Phelps2013; Raw et al., Reference Raw, Riddin, Wasserman, Lehman, Bornman and Adams2020; Cahoon et al., Reference Cahoon, McKee and Morris2021; Saintilan et al., Reference Saintilan, Kovalenko, Guntenspergen, Rogers, Lynch, Cahoon, Lovelock, Friess, Ashe and Krauss2022). The effect of sea-level rise on the position in the tidal frame of BCEs has been conceptualised by Allen (Reference Allen2000) to account for the addition of mineral and organic material, autocompaction and relative sea-level rise. Providing accommodation is available, below-ground biomass from established vegetation increases substrate volume and the mass of the soil organic carbon pool; above-ground biomass baffles tidal energy, improving hydrodynamic conditions for the deposition of suspended sediments. As substrate elevations increase, the space available for below-ground organic matter additions and mineral sediment addition diminishes, and organic matter decomposition may increase due to an associated reduction in inundation depth, duration and/or frequency (often termed hydroperiod). In combination, these factors generate a self-organising negative feedback that favours the stabilisation of substrate elevations. As a small increment in sea level increases accommodation, the associated increase in tidal inundation serves to enhance conditions favouring the accumulation of mineral and organic sediments, thereby offsetting the small increment in sea level, and maintaining the intertidal position of the substrate; addition of organic material is a vital component of this negative feedback. The coupling between inundation and organic matter addition that contributes to this negative feedback was initially conceptualised for marshes of the SE coast of the USA (Morris et al., Reference Morris, Sundareshwar, Nietch, Kjerfve and Cahoon2002) and has formed the basis for models projecting the organic response of substrates to sea-level rise (Mudd et al., Reference Mudd, Howell and Morris2009; Mack et al., Reference Mack, Lane, Deng, Morris and Bauer2023). Field studies have also confirmed linkages between coastal wetland evolution, accommodation space and carbon concentrations in mangrove and saltmarsh substrates of SE Australia (Owers et al., Reference Owers, Woodroffe, Mazumder and Rogers2022) (Figure 3).

Figure 3. (A) Conceptual model of lateral zonation of BCEs of southeastern Australia with respect to tidal parameters, and varying distribution of soil organic carbon within the active root zone, inactive root zone and subtidal zone; and (B) associated generalised variation in carbon storage within BCEs (modified from Owers et al. (Reference Owers, Woodroffe, Mazumder and Rogers2022)). (C and D) Relationships between carbon storage and sea-level change (from T 0 to T n) under conditions of (C) relatively stable sea level since the mid-Holocene and (D) rising sea level since the mid-Holocene (modified from Allen (Reference Allen2000)).

The PAST: Sea-level rise, and blue carbon accommodation

For extended periods over Earth’s history, when the coincidence of rising sea level, favourable coastal geomorphology and suitable tidal range has been conducive to extensive coastal wetland development, blue carbon has been an important and arguably dominant control on global trends in atmospheric CO2. During the Oligo-Miocene, the combined influence of sea-level rise, high tidal range and a resultant extensive mangrove development in the South China Sea trapped up to 2000 Pg of organic carbon, equivalent to up to 60 p.p.m. of atmospheric CO2 per Myr (Collins et al., Reference Collins, Avdis, Allison, Johnson, Hill, Piggott, Hassan and Damit2017). The development of these forests could have been a major contributor to the reduction in atmospheric CO2 concentrations from circa 800 to 300 p.p.m. since the Late Oligocene (34-0 Ma) (Collins et al., Reference Collins, Avdis, Allison, Johnson, Hill, Piggott, Hassan and Damit2017).

At the peak of the last glacial maximum, sea level was 130–120 m lower than present; this low stand and the present high stand are indicative end points of global eustatic sea-level cycles (Murray-Wallace and Woodroffe, Reference Murray-Wallace and Woodroffe2014). The response of mangrove forests and saltmarshes to sea-level rise since the last glacial maximum has been likened to the behaviour of coral reefs at the same period (Neumann and MacIntyre, Reference Neumann, Macintyre, Gabrie, Toffart and Salvat1985; Reed, Reference Reed1990; Woodroffe and Davies, Reference Woodroffe, Davies, Perillo, Wolanski, Cahoon and Brinson2009), where, depending upon the rate of sediment supply relative to the rate of sea-level rise, mangrove forests and saltmarshes may be ‘drowned’, ‘backstep’, ‘catch-up’, ‘keep-up’, ‘prograde’ or ‘emerge’. When rates of sea-level rise exceeded 1 m per century during the late-Pleistocene and early-Holocene, the capacity of mangrove forests to accumulate mineral and organic material appears to have been exhausted, and evidence of ‘drowned’ mangrove peats overtopped by marine transgressive sand sheets have been preserved on the Sahul Shelf (northwest Western Australia (Nicholas et al., Reference Nicholas, Nichol, Howard, Picard, Dulfer, Radke, Carroll, Tran and Siwabessy2014) and the Sunda Shelf (on the western rim of the South China Sea) at water depths of up to 100 m (Hanebuth et al., Reference Hanebuth, Stattegger and Grootes2000). Extensive mangrove forest development at the time appears to have been terminated by Meltwater Pulse 1A, during which rates of relative sea-level rise increased to >20 mm yr.−1 (Lambeck et al., Reference Lambeck, Rouby, Purcell, Sun and Sambridge2014).

Deceleration in the rate of sea-level rise in the early-Holocene was broadly marked by mangrove development in the millennia prior to sea-level stabilisation near present levels. In settings of high sediment yield, including the Ganges-Brahmaputra delta, India (Hait and Behling, Reference Hait and Behling2009) and the Queensland continental shelf, Australia (Grindrod et al., Reference Grindrod, Moss and Kaars1999), mangrove forests adjusted in situ to sea-level rise from the early-Holocene (~9,000 BP), but were subsequently ‘drowned’ and then ‘backstepped’. By ~7,500 BP, relative sea-level rise had decelerated to less than 6–7 mm yr.−1, and ‘catch-up’, or mangrove landward transgression, followed by ‘progradation’ occurred where sediment supply was high, resulting in extensive mangrove forests in tropical minerogenic settings. In many places, these forests were considerably greater in extent than contemporary mangrove forests, including Australia (Woodroffe et al., Reference Woodroffe, Thom and Chappell1985), the Mekong and Red River deltas of Vietnam (Tran and Ngo, Reference Tran, Ngo and Nguyen2000; Li et al., Reference Li, Saito, Mao, Tamura, Song, Zhang, Lu, Sieng and Li2012), and the Great Songkla Lakes, Thailand (Horton et al., Reference Horton, Gibbard, Mine, Morley, Purintavaragul and Stargardt2005). High rates of organic matter accumulation in this globally synchronous phase of blue carbon development sequestered an estimated 20–60 Pg. C, contributing to a 5 p.p.m. decline in atmospheric CO2 concentrations in the early Holocene (Saintilan et al., Reference Saintilan, Khan, Ashe, Kelleway, Rogers, Woodroffe and Horton2020). An early-mid Holocene decline in methane, primarily in the Southern Hemisphere, according to ice-core data (Beck et al., Reference Beck, Bock, Schmitt, Seth, Blunier and Fischer2018), commenced over the same period, with reductions in Southern Hemisphere emissions estimates of ~19 Tg CH4 yr.−1 (Beck et al., Reference Beck, Bock, Schmitt, Seth, Blunier and Fischer2018). The δ13C signals in methane in the Southern Hemisphere for the period show a 1.5 p.p.t. depletion (Beck et al., Reference Beck, Bock, Schmitt, Seth, Blunier and Fischer2018) consistent with a replacement of vegetation utilising the C4 photosynthetic pathway (tropical grasslands and saltmarsh adapted to low atmospheric carbon dioxide) with mangroves utilising the C3 pathway (Sowers, Reference Sowers2010).

By the mid-Holocene, eustatic sea level stabilised within approximately 2 m of its current elevation (Clark et al., Reference Clark, Farrell and Peltier1978; Khan et al., Reference Khan, Ashe, Shaw, Vacchi, Walker, Peltier, Kopp and Horton2015). However, the initiation of mangrove and saltmarsh transgressive phases was globally variable due to the influence of glacio-isostatic adjustment on varying rates of mid- to late-Holocene relative sea-level rise (Ribeiro et al., Reference Ribeiro, Batista, Cohen, França, Pessenda, Fontes, Alves and Bendassolli2018), and the modulating effect of other climatic variables (e.g. droughts and/or frequent storms) on conditions conducive to intertidal vegetation expansion or decline (Sherrod and McMillan, Reference Sherrod and McMillan1985; Jones et al., Reference Jones, Wingard, Stackhouse, Keller, Willard, Marot, Landacre and Bernhardt2019). Global scale variation in relative sea-level trends, largely arising from glacio-isostatic adjustment, had a profound influence on blue carbon accumulation throughout the mid- to late-Holocene and storage since this time (Rogers et al., Reference Rogers, Kelleway, Saintilan, Megonigal, Adams, Holmquist, Lu, Schile-Beers, Zawadzki, Mazumder and Woodroffe2019a; Reference Rogers, Macreadie, Kelleway and Saintilan2019b). Principally related to distance from regions of maximal ice sheet extent during the last glacial period, relative sea-level rise modifies the accommodation available for blue carbon. The delineation of zones across oceans globally where post-glacial sea-level trends are relatively similar (Figure 4A and 4B, Clark et al., Reference Clark, Farrell and Peltier1978) therefore provides an indication of the accommodation available for blue carbon storage over the past few millennia. While more is known of post-glacial sea-level change since these zones were initially demarcated, their broad correspondence with what we now know of relative sea-level trends (Khan et al., Reference Khan, Ashe, Shaw, Vacchi, Walker, Peltier, Kopp and Horton2015) provides some confidence in the geographic position of zones (noting that boundaries between zones are diffuse and not definite). Clark et al. (Reference Clark, Farrell and Peltier1978) delineated five zones across oceans globally (a sixth zone was associated with continental coastlines) that can be broadly grouped into three regions: (i) near-field locations are proximal to ice sheets of the last glacial maximum and typically exhibit continuous patterns of relative sea-level fall (i.e. Clark et al., Reference Clark, Farrell and Peltier1978, zone I); (ii) intermediate locations exhibit complex sea-level trends; however, relative sea-level rise over the past few millennia is typical; and (iii) far-field locations are distal from ice sheets and eustatic sea-level trends dominate processes of glacio-isostatic adjustment. For brevity, we focus on end members: far-field locations (Zones IV and V) where the relative sea level has been relatively stable for millennia (or may have fallen); intermediate locations where relative sea level has been rising over the mid- to late-Holocene (Zones II and III); and near-field locations (Zone I) where relative sea level has been falling (Figure 4A).

Figure 4. Relative sea-level change is a significant control on processes of carbon accumulation and decomposition, and varies globally according to the generalised Holocene relative sea-level zones (A) (Clark et al., Reference Clark, Farrell and Peltier1978) and generalised Holocene relative sea-level curves across these zones (B). Note the distribution of coastal acid sulphate soils in (Michael, Reference Michael2013) (A), which corresponds broadly with regions where sea-level conditions facilitated widescale mangrove and saltmarsh development throughout the late-Holocene. When sea levels are rising (C), sedimentary carbon continues to accumulate within available accommodation and pathways of decomposition are dampened under increasingly anaerobic conditions. Where sea level has been relatively stable (D), blue carbon additionality is limited by the upper limit of tidal inundation and substrates become increasingly mineral dominated and support terrestrial vegetation as accommodation diminishes. Under conditions of falling sea levels (E), substrates become increasingly terrestrialised (i.e. with terrestrial vegetation) and conditions favour aerobic decomposition and methanogenesis of blue carbon.

Far-field locations (Zones IV and V), distal from regions of maximal ice sheet extent, exhibit patterns of mid-Holocene infill in mangrove (Woodroffe et al., Reference Woodroffe, Mulrennan and Chappell1993; Cohen et al., Reference Cohen, Filho, Lara, Behling and Angulo2005; Hashimoto et al., Reference Hashimoto, Saintilan and Haberle2006; Proske and Haberle, Reference Proske and Haberle2012; França et al., Reference França, Cohen, Pessenda, Rossetti, Lorente, Junior, Guimarães, Friaes and Macario2013; Boski et al., Reference Boski, Bezerra, de Fátima, Souza, Maia and Lima-Filho2015; Punwong et al., Reference Punwong, Selby and Marchant2018) and saltmarsh settings, with coastal barriers typically enclosing bays along the more southern wave-dominated coastlines (Compton, Reference Compton2001; Vilanova et al., Reference Vilanova, Prieto and Espinosa2006; Fornari et al., Reference Fornari, Giannini and Junior2012; Kennedy et al., Reference Kennedy, Wong and Jacobsen2021). Ongoing coastal and estuarine sedimentation, and a fall in sea level to present levels where a high stand occurred in the late Holocene, caused coastal floodplains to increasingly ‘prograde’ or become ‘emergent’, with accommodation being limited, and BCEs restricted to the fringes of estuarine shorelines (Woodroffe and Davies, Reference Woodroffe, Davies, Perillo, Wolanski, Cahoon and Brinson2009). BCEs were replaced by floodplain terrestrial forests and freshwater wetlands, a shift that may have contributed to gradual increases in atmospheric methane concentrations in the late-Holocene after declining in the early- to mid-Holocene (Beck et al., Reference Beck, Bock, Schmitt, Seth, Blunier and Fischer2018). Preservation of soil organic carbon in far-field locations is limited by decomposition as coastal floodplains become increasingly terrestrialised and support grasses and sedges, else soil organic matter may be vulnerable to metabolisation and formation of pyrites when sulphate and sulphate reducing bacteria are present. This sets up the conditions for generation of acid sulphate soils when coastal floodplains are drained, and oxidisation occurs. The relationship between former distribution of saltmarshes and mangrove forests, relative sea-level stability and coastal acid sulphate soil development is well established (Pons et al., Reference Pons, Van Breemen and Driessen1982; Van Breemen, Reference Van Breemen1982) and reflected in the greater extent of actual and potential coastal acid sulphate soils in SE Asia, Africa, Australia and South America (Michael, Reference Michael2013) and their virtual absence from intermediate-field locations (see Figure 4A).

Intermediate field locations also exhibit broad agreement in geomorphological evolution over the Holocene (Woodroffe, Reference Woodroffe1981; Digerfeldt and Hendry, Reference Digerfeldt and Hendry1987; Parkinson, Reference Parkinson1989; Parkinson et al., Reference Parkinson, Ron and White1994; McKee, Reference McKee2011). In Florida and the northern Gulf of Mexico, the early Holocene was marked by rates of relative sea-level rise that were too high for broadscale mangrove development (<~7,500 BP) (Sherrod and McMillan, Reference Sherrod and McMillan1985; Parkinson, Reference Parkinson1989). Evident from interbedded peats and marls, the onset of the transgressive ‘catch-up’ phase occurred from about 3,500 BP (Scholl, Reference Scholl1964; Parkinson et al., Reference Parkinson, Ron and White1994; Jones et al., Reference Jones, Wingard, Stackhouse, Keller, Willard, Marot, Landacre and Bernhardt2019), although the occurrence of continuous vertical peat growth, typical of ‘keep-up’ behaviour, is modulated in some locations by other climatic factors, including a period of cooling, that may not have been conducive to widespread mangrove expansion and vertical growth (Sherrod and McMillan, Reference Sherrod and McMillan1985; Jones et al., Reference Jones, Wingard, Stackhouse, Keller, Willard, Marot, Landacre and Bernhardt2019). Where conditions were favourable, vertical growth of mangrove peats is near continuous. In particular, the cenotes of the Yucatan Peninsula are reported to have among the highest mangrove carbon stocks globally, and their accumulation has been related to ongoing relative sea-level rise over the late-Holocene (Adame et al., Reference Adame, Santini, Torres-Talamante and Rogers2021). The later onset of transgressive phases of mangrove development in intermediate field locations reflect stronger rates of relative sea-level rise throughout the mid-Holocene, ongoing sea-level rise in the late-Holocene, the influence of other climatic variables, as well as limited capacity for vertically adjustment to sea-level rise due to low rates of sediment supply in some carbonate dominated settings. Accordingly, ongoing sea-level rise and limited mineral sediment supply since the late-Holocene may explain the preservation of carbon-rich mangrove peats in this region (McKee, Reference McKee2011). Similar preservation of saltmarsh peats and associated foraminifera is evident at intermediate-field sites with a history of increasing accommodation with sea-level rise throughout the mid- to late-Holocene (Redfield, Reference Redfield1972; Orson et al., Reference Orson, Warren and Niering1998; Gehrels, Reference Gehrels1999).

Saltmarshes at near-field locations (climate not suitable for mangroves throughout the Holocene), those proximal to regions of maximum ice sheet extent at the last glacial maximum, exhibited a highly variable pattern of vertical growth and carbon accumulation dependent upon the influence of glacio-isostatic adjustment on relative sea-level rise (Khan et al., Reference Khan, Ashe, Shaw, Vacchi, Walker, Peltier, Kopp and Horton2015). Analyses of radiocarbon dated saltmarsh sequences in the UK differentiated both transgressive sequences (i.e. ‘catch-up’), representing increasing marine influence and regressive sequences (i.e. ‘progradation’ or ‘emergent’) indicating increasing terrestrialisation (Horton et al., Reference Horton, Shennan, Bradley, Cahill, Kirwan, Kopp and Shaw2018). The presence of these sequences aligned with Holocene sea-level history, with ‘catch-up’ transgressive sequences predominating in southern England where relative sea-level rise exhibited a pattern of deceleration, while regressive sequences were more prominent in Scotland, where relative sea level fell in both the early- and late-Holocene. Complex spatio-temporal patterns of postglacial relative sea-level change throughout the mid- to late-Holocene are also preserved in saltmarsh peats of near-field locations across the coastline of the North Atlantic (Vacchi et al., Reference Vacchi, Engelhart, Nikitina, Ashe, Peltier, Roy, Kopp and Horton2018; Cohen et al., Reference Cohen, Cartelle, Barnett, Busschers and Barlow2022; Creel et al., Reference Creel, Austermann, Khan, D’Andrea, Balascio, Dyer, Ashe and Menke2022).

Patterns of relative sea-level change over the Holocene and its influence on the distribution of BCEs have provided important lines of evidence about the future of BCEs. Global analyses of soil organic carbon stocks in saltmarshes indicate that regions where sea level has a longer history of rising over the mid- to late-Holocene, that is intermediate and some near-field locations, exhibit higher soil organic carbon stocks and deeper soil organic carbon pools than far-field locations (Rogers et al., Reference Rogers, Kelleway, Saintilan, Megonigal, Adams, Holmquist, Lu, Schile-Beers, Zawadzki, Mazumder and Woodroffe2019a). This has been linked to the influence of relative sea-level rise on accommodation for blue carbon within substrates, the largest carbon pool within BCEs. Where sea level has been rising at a moderate rate for a few millennia, vertical space is created for storage of blue carbon within below-ground biomass and soils (Figure 4C). In contrast, where sea level has been relatively stable over the mid- to late-Holocene, that is across much of the Southern Hemisphere, carbon pools may be depleted and shallower (although the depth is dependent upon tidal range), and this has been related to the limitations placed on accommodation as substrates become increasingly dominated by mineral sediments (Figure 4D). Where sea level has been falling, substrates become increasingly terrestrialised with brackish to fresh substrate salinities that are aerobic and favour decomposition and methanogenesis (Figure 4E). These differences in organic matter content between far-, intermediate- and near-field locations are also reflected in the character of contemporary sediment accumulating in wetlands. Intermediate- and near-field locations have higher organic carbon accumulation above artificial marker horizons (Saintilan et al., Reference Saintilan, Khan, Ashe, Kelleway, Rogers, Woodroffe and Horton2020), reflecting both the comparatively higher contemporary rates sea-level rise, and possibly the more organic sub-tidal reservoirs of sediment contributing to marsh accretion (Hopkinson et al., Reference Hopkinson, Morris, Fagherazzi, Wollheim and Raymond2018).

The FUTURE: The present and past as a guide to blue carbon futures

The spectre of climate change has focussed attention on reducing atmospheric carbon concentrations and limiting warming well below 2 °C, a commitment established in the Paris Agreement at COP21 (Iyer et al., Reference Iyer, Edmonds, Fawcett, Hultman, Alsalam, Asrar, Calvin, Clarke, Creason and Jeong2015). Accordingly, the capacity of BCEs to draw carbon from the atmosphere is being leveraged as a climate mitigation strategy, and received considerable recognition in the latest IPCC report (Cooley et al., Reference Cooley, Schoeman, Bopp, Boyd, Donner, Ghebrehiwet, Ito, Kiessling, Martinetto, Ojea, Racault, Rost, Skern-Mauritzen, Pörtner, Roberts, Tignor, Poloczanska, Mintenbeck, Alegría, Craig, Langsdorf, Löschke, Möller, Okem and Rama2022). Conservation of BCEs aimed at minimising losses in extent through land use and land cover change (LULCC) and restoring condition through improved management will be important. The Reducing Emissions from Deforestation and Forest Degradation in Developing Countries (REDD+) program of the United Nations Framework Convention on Climate Change (UNFCCC) specifically targets the improved management of forests to minimise loss and release of greenhouse gases, and conservation efforts are increasing in developing countries (Ahmed and Glaser, Reference Ahmed and Glaser2016). Similar programs for non-forested ecosystems, such as saltmarshes, do not exist; however, there is a global peatland initiative that could be applied to saltmarshes. In many jurisdictions BCEs are already protected from loss because of the benefits they provide to society (Romañach et al., Reference Romañach, DeAngelis, Koh, Li, Teh, Barizan and Zhai2018). Despite these policies, loss of BCEs is ongoing (Gedan et al., Reference Gedan, Silliman and Bertness2009; Friess et al., Reference Friess, Rogers, Lovelock, Krauss, Hamilton, Lee, Lucas, Primavera, Rajkaran and Shi2019; Goldberg et al., Reference Goldberg, Lagomasino, Thomas and Fatoyinbo2020), and this increases the burden for carbon additionality by other mechanisms, such as restoration.

Restoring BCEs, achieved by planting vegetation, seeds or propagules, or managing barriers to tidal exchange that have been put in place to facilitate past LULCC, may enhance carbon sequestration as blue carbon vegetation re-establishes. Analyses suggest that restoration to recover BCE habitat that has been lost due to human activities in the coastal zone is potentially feasible for mangroves, less so for seagrasses and saltmarshes (Griscom et al., Reference Griscom, Adams, Ellis, Houghton, Lomax, Miteva, Schlesinger, Shoch, Siikamäki, Smith, Woodbury, Zganjar, Blackman, Campari, Conant, Delgado, Elias, Gopalakrishna, Hamsik, Herrero, Kiesecker, Landis, Laestadius, Leavitt, Minnemeyer, Polasky, Potapov, Putz, Sanderman, Silvius, Wollenberg and Fargione2017; Macreadie et al., Reference Macreadie, Costa, Atwood, Friess, Kelleway, Kennedy, Lovelock, Serrano and Duarte2021). However, the capacity for large-scale restoration is constrained by socio-economic factors, particularly where the coastal zone is critical for maintaining livelihoods and food security (Herr et al., Reference Herr, Blum, Himes-Cornell and Sutton-Grier2019). Efforts are in place globally to restore mangrove forests (Friess et al., Reference Friess, Rogers, Lovelock, Krauss, Hamilton, Lee, Lucas, Primavera, Rajkaran and Shi2019), but success is highly variable; often there is a lack of understanding of the geomorphological and hydrological controls on restoration success, and mangrove restoration efforts have been plagued with failure (Lee et al., Reference Lee, Hamilton, Barbier, Primavera and Lewis2019; Lovelock et al., Reference Lovelock, Barbier and Duarte2022c). Saltmarsh restoration is also occurring, but receives considerably less scientific attention beyond North America; this likely reflects an efficient policy environment and sufficient financial capacity for restoration (Billah et al., Reference Billah, Bhuiyan, Islam, Das and Hoque2022). Despite these challenges, restoration of BCEs is likely to accelerate, being buoyed by the United Nations declaration that 2021–2030 is the “United Nations Decade on Ecosystem Restoration” to help meet sustainable development goals (Billah et al., Reference Billah, Bhuiyan, Islam, Das and Hoque2022).

Market mechanisms have been developed to incentivise BCE restoration, and currently there are two primary markets; the compliance and voluntary markets (Sapkota and White, Reference Sapkota and White2020). Compliance markets are underpinned by regulations to offset greenhouse gas emissions and typically require offsets to be accounted for under existing standards and using approved methodologies, such as the Verified Carbon Standard (VCS) methodology for tidal wetlands (VM0033) (Emmer et al., Reference Emmer, Needelman, Emmett-Mattox, Crooks, Megonigal, Myers, Oreska, McGlathery, Shoch and Washington2015a; Emmer et al., Reference Emmer, von Unger, Needelman, Crooks and Emmett-Mattox2015b). In some jurisdictions, the voluntary market is also highly regulated; for example, the Australian Government administers a voluntary market, the Emissions Reduction Fund, which provides tradeable credits (Australian Carbon Credit Units) for tidal restoration activities that increase blue carbon storage (Lovelock et al., Reference Lovelock, Adame, Bradley, Dittmann, Hagger, Hickey, Hutley, Jones, Kelleway and Lavery2022a; Lovelock et al., Reference Lovelock, Adame, Butler, Kelleway, Dittmann, Fest, King, Macreadie, Mitchell and Newnham2022b). The payment period for these programs is well-defined as the carbon benefits are likely to diminish in the above-ground carbon pool once vegetation has reached maturity, and in the below-ground pool when substrates are saturated with mineral and organic material (i.e. when accommodation is limited). The 25-year permanence time frame aligns with the period for which woody vegetation is anticipated to reach maturity and exhibit high rates of carbon addition to substrates (Osland et al., Reference Osland, Feher, Spivak, Nestlerode, Almario, Cormier, From, Krauss, Russell and Alvarez2020), while the 100-year timeframe aligns with what is regarded to be permanently sequestered soil organic carbon (i.e. permanence) (Dynarski et al., Reference Dynarski, Bossio and Scow2020). Payments are dependent upon forward projections and ongoing verification.

Use of BCEs as a mechanism for carbon removal has received some criticism (Williamson and Gattuso, Reference Williamson and Gattuso2022) due to the high variability and errors in carbon burial rates, lateral carbon transport (Maher et al., Reference Maher, Call, Santos and Sanders2018), methane and nitrous oxide fluxes (Rosentreter et al., Reference Rosentreter, Al‐Haj, Fulweiler and Williamson2021a; Malerba et al., Reference Malerba, Friess, Peacock, Grinham, Taillardat, Rosentreter, Webb, Iram, Al-Haj and Macreadie2022), carbonate formation and dissolution (Saderne et al., Reference Saderne, Geraldi, Macreadie, Maher, Middelburg, Serrano, Almahasheer, Arias-Ortiz, Cusack and BDJNc2019; Van Dam et al., Reference Van Dam, Zeller, Lopes, Smyth, Böttcher, Osburn, Zimmerman, Pröfrock, Fourqurean and Thomas2021), vulnerability to future climate change and non-climatic factors, and cost-effectiveness and scaleability (Macreadie et al., Reference Macreadie, Costa, Atwood, Friess, Kelleway, Kennedy, Lovelock, Serrano and Duarte2021). Confidence in forward projections is likely to be improved as data collection continues and knowledge gaps are addressed (Macreadie et al., Reference Macreadie, Anton, Raven, Beaumont, Connolly, Friess, Kelleway, Kennedy, Kuwae, Lavery, Lovelock, Smale, Apostolaki, Atwood, Baldock, Bianchi, Chmura, Eyre, Fourqurean, Hall-Spencer, Huxham, Hendriks, Krause-Jensen, Laffoley, Luisetti, Marbà, Masque, KJ, Megonigal, Murdiyarso, Russell, Santos, Serrano, Silliman, Watanabe and Duarte2019).

Permanence (i.e. beyond the 25-year and 100-year time frames) of blue carbon is fundamental to the success of BCEs as a natural climate solution. Critically, conservation and restoration activities will not occur in the absence of sea-level rise, warming and elevated atmospheric carbon dioxide concentrations, placing considerable uncertainty regarding permanence. The fate of sequestered carbon from BCEs once they succumb to sea-level rise is difficult to project, but likely to be highly variable. Palaeo-records provide the opportunity to validate projections of the response of BCEs to sea-level rise; however, do not fully indicate blue carbon futures as sea-level rise is now occurring on coastal landscapes that developed throughout the Holocene, have been highly modified and do not have a historic analogue (Woodroffe and Murray-Wallace, Reference Woodroffe and Murray-Wallace2012). Projecting blue carbon futures therefore requires integration of information from the past and present behaviour.

Recent analyses have indicated that rising seas associated with ice melt following the last glacial maximum exceeded the capacity of tropical mangroves (Saintilan et al., Reference Saintilan, Khan, Ashe, Kelleway, Rogers, Woodroffe and Horton2020) and saltmarshes (Horton et al., Reference Horton, Shennan, Bradley, Cahill, Kirwan, Kopp and Shaw2018; Törnqvist et al., Reference Törnqvist, Cahoon, Morris and Day2021) to remain in situ (i.e. ‘keep-up’) when sea level increased at rates exceeding ~5–7 mm yr.−1. However, mangrove and saltmarsh sediments have been preserved since the Holocene (Hanebuth et al., Reference Hanebuth, Stattegger and Grootes2000; McKee et al., Reference McKee, Cahoon and Feller2007; Wang et al., Reference Wang, Sun, Wang and Stattegger2009) following sea-level rise at rates higher than currently encountered (Redfield, Reference Redfield1972; McKee et al., Reference McKee, Cahoon and Feller2007; Saintilan et al., Reference Saintilan, Khan, Ashe, Kelleway, Rogers, Woodroffe and Horton2020). Long-periods of sea-level stability across much of the Southern Hemisphere have contributed to the development of broad, mature coastal floodplains (i.e. considerable elevation capital) when conditions are conducive. These floodplains are not typically saturated with carbon (Rogers et al., Reference Rogers, Kelleway, Saintilan, Megonigal, Adams, Holmquist, Lu, Schile-Beers, Zawadzki, Mazumder and Woodroffe2019a), may be hot spots for potential and realised acid sulphate soils (Michael, Reference Michael2013), and could become extensive BCEs, much like those of the mid-Holocene (Woodroffe et al., Reference Woodroffe, Thom and Chappell1985). Many BCEs in the northern hemisphere have been adapting to sea-level rise for millennia and are likely to be lower-lying (i.e. less elevation capital) and have limited capacity to adapt to anticipated sea-level rise as the coastal zone is highly contested and coastal squeeze is likely.

The permanence of organic material exposed to drowning is an important consideration. Since it takes some time for mature trees and tidal marshes, that are high in the tidal frame to drown with incremental increases in sea level, it is feasible that submergence and death may occur beyond the minimum 25-year permanency time frame applied in managed blue carbon markets. Loss of standing biomass that is currently lower in the tidal frame or in the interior of saltmarshes may be high (Kearney et al., Reference Kearney, Stevenson and Ward1994), and depending on exposure to erosion, soil carbon may be variably preserved (Krauss et al., Reference Krauss, Demopoulos, Cormier, From, McClain-Counts and Lewis2018; Rogers et al., Reference Rogers, Kelleway, Saintilan, Megonigal, Adams, Holmquist, Lu, Schile-Beers, Zawadzki, Mazumder and Woodroffe2019a) or reworked and transported elsewhere (DeLaune and White, Reference DeLaune and White2012; Haywood et al., Reference Haywood, Hayes, White and Cook2020). Increasing atmospheric carbon dioxide concentrations may enhance productivity of BCE vegetation, with enhanced root allocation contributing to sea-level rise adaptation and carbon sequestration via addition of carbon volume to substrates (Ball et al., Reference Ball, Cochrane and Rawson1997; Langley et al., Reference Langley, McKee, Cahoon, Cherry and Megonigal2009; McKee et al., Reference McKee, Rogers, Saintilan and Middleton2012; Reef et al., Reference Reef, Spencer, Mӧller, Lovelock, Christie, McIvor, Evans and Tempest2017). Widespread mangrove dieback has been associated with short-term methane flux to the atmosphere (Jeffrey et al., Reference Jeffrey, Reithmaier, Sippo, Johnston, Tait, Harada and Maher2019); this has potential implications on atmospheric carbon budgets in the event of broadscale mangrove mortality under high rates of sea-level rise, or short-term sea-level fluctuations. Figure 5 conceptualises the hypothesised carbon storage and greenhouse gas flux outcomes under scenarios of atmospheric carbon dioxide concentrations, warming and relative sea-level rise.

Figure 5. Conceptualisation of the interacting effects of atmospheric carbon dioxide, warming and relative sea-level rise on BCE projected to occur under a range of emissions scenarios. Under the baseline scenario (A) carbon is fixed by in situ vegetation and contributes to soil carbon accumulation and substrate volume via accretion. The landward margin under brackish conditions is a source of methane. Under the mid-range emissions scenario (B) the feedbacks between elevated atmospheric CO2 and organic carbon sequestration and between relative sea-level rise (RSLR) and organic carbon production, preservation, and vertical accretion are strengthened. Saline intrusion reduces methane emissions in the landward fringe, although this may be counterbalanced by increased emissions resulting from increases in NPP induced by CO2 and warming. Under the high-range emission scenario (C) a tipping point is reached where RSLR exceeds vertical accretion leading to mortality and shoreline retreat. Mortality of terrestrial vegetation contributes to elevated methane production in the short term. Note that the relative strength of interactions with greenhouse gases is indicated by the thickness of lines, and greyed-out vegetation is indicative of dieback or loss associated with the effects of rising sea levels.

The degree to which climate change modifies BCEs is difficult to project as shoreline erosion is poorly preserved in the stratigraphy of depositional environments, limiting the capacity to parameterise models. In addition, multiple coastal processes contribute to shoreline change and observations indicate considerable local and regional variability in the operation of processes. Studies that project BCE dynamics with sea-level rise and infer carbon implications are therefore typically undertaken at the local scale, and at data-rich sites, and the outcomes can rarely be extrapolated to other locations. Otherwise, projections are dependent upon simplification of processes and apply reductive estimates of carbon concentrations (Lovelock and Reef, Reference Lovelock and Reef2020; Wang et al., Reference Wang, Sanders, Santos, Tang, Schuerch, Kirwan, Kopp, Zhu, Li and Yuan2021), or explore future scenarios using simplified idealised models (Kirwan and Mudd, Reference Kirwan and Mudd2012). Spatial models typically require decisions regarding whether landward retreat of BCEs is parameterised, and, invariably, the projected outcomes are highly dependent upon model parameterisation. For example, recent projections estimated net gains in blue carbon in the order of 1.5 Pg to 2100 when coastal squeeze impacts are minimised and climate change impacts are high (i.e. RCP8.5 scenario), while net blue carbon gains are in the order of 0.8 Pg to the year 2100 when under moderate climate change scenario (i.e. RCP4.5 scenario) (Lovelock and Reef, Reference Lovelock and Reef2020).

Models have consistently indicated the importance of minimising coastal squeeze to enhance climate adaptation and mitigation benefits from BCEs (Schuerch et al., Reference Schuerch, Spencer, Temmerman, Kirwan, Wolff, Lincke, McOwen, Pickering, Reef, Vafeidis, Hinkel, Nicholls and Brown2018; Lovelock and Reef, Reference Lovelock and Reef2020; Wang et al., Reference Wang, Sanders, Santos, Tang, Schuerch, Kirwan, Kopp, Zhu, Li and Yuan2021). Managing structures that modify tidal exchange and sediment supply will be crucial for maximising blue carbon benefits. Indeed, storm surge barriers are already managed to reduce coastal flooding impacts (Haigh et al., Reference Haigh, Dornbusch, Brown, Lyddon, Nicholls, Penning-Roswell and Sayers2022) and engineering structures could be modified to meet design requirements anticipated with sea-level rise and to manage for blue carbon services (Sadat-Noori et al., Reference Sadat-Noori, Rankin, Rayner, Heimhuber, Gaston, Drummond, Chalmers, Khojasteh and Glamore2021; Haigh et al., Reference Haigh, Dornbusch, Brown, Lyddon, Nicholls, Penning-Roswell and Sayers2022). In many cases ‘holding back the tide’ will become challenging and costly, and it is likely that difficult land use decisions will be made to increase BCE extent and the provision of co-benefits (Rogers et al., Reference Rogers, Lal, Asbridge and Dwyer2022). Sea-level rise will reduce the viability of large tracts of low-lying coastal land for agriculture and grazing purposes, and the efficacy of existing structures to hold back tides, many of which were designed when rates of sea-level rise were negligible, will be tested. In these circumstances, the benefits of land cover conversion for BCEs should therefore be weighed against the costs associated with upgrading existing structures to meet future design requirements that account for the effects of sea-level rise and storm surges on tidal regimes. Momentum towards recognising BCE co-benefits for biodiversity, coastal fisheries and water quality is increasing (Rahman et al., Reference Rahman, Zimmer, Ahmed, Donato, Kanzaki and Xu2021; Hagger et al., Reference Hagger, Waltham and Lovelock2022), and efforts are underway to develop a ‘blue chip’ carbon markets that provide payments for blue carbon additionality and co-benefits (Macreadie et al., Reference Macreadie, Robertson, Spinks, Adams, Atchison, Bell-James, Bryan, Chu, Filbee-Dexter and Drake2022). These blue-chip markets may sufficiently incentivise land managers to reconsider upgrades of tidal barriers and instead receive a blue carbon income as the land is restored to a BCE (Rogers et al., Reference Rogers, Lal, Asbridge and Dwyer2022).

Conclusion

In the short- to medium-term, climate change may increase the capacity of BCEs to capture and store atmospheric CO2, largely due to processes that respond to elevated CO2 and temperature, and influence carbon capture and storage. These processes include in situ responses such as CO2 fixation, biomass storage, biogeochemical enhancement of burial efficiency, as well as the expansion of BCEs at local and global scales. The magnitude and duration of the negative feedback on climate may vary between hemispheres. Climate change is expected to squeeze mangrove and saltmarsh areas in the Northern Hemisphere between accelerating relative sea-level rise and hard barriers. In the Southern Hemisphere, opportunities for landward expansion of mangrove and saltmarsh may be available where late-Holocene sea-level history coupled with ongoing sediment supply and lower contemporary rates of relative sea-level rise has facilitated the development of broad coastal floodplains and where coastal squeeze effects are minimised. The long-term future for these negative feedbacks on radiative forcing is dependent upon decisions made in the coming decades. At the global scale, the rate of sea-level rise projected under high emissions scenarios will lead to degradation and loss of existing coastal wetlands. Sea-level rise of ~5–7 mm yr.−1 is likely to be a critical tipping point at which the predominantly negative climate feedbacks driven by blue carbon sequestration become positive feedbacks driven by plant decomposition and remineralisation. This tipping point will be surpassed under high emissions scenarios within the next century.

The threat to in situ coastal wetlands makes local land-use and coastal protection the key determinant of long-term survival, driven by retreat of BCEs to higher elevations. Investment in coastal wetland conservation and restoration provides benefits not only for the preservation of ecosystem services such as coastal fisheries, but also a promising opportunity for nature-based mitigation. National governments are developing a broad spectrum of climate adaptation and mitigation responses with innovative approaches to financing these activities, including some focused specifically on blue carbon. Carbon markets are rapidly expanding as a tool for governments, private corporations, and individuals to reduce greenhouse gas emissions. While currently only a few blue carbon projects have reached the point of generating finance through carbon markets, projects are in development in Mexico, Kenya, Colombia, Madagascar and other locations, and blue carbon credits are in high demand due to the multitude of co-benefits provided. This suggests that carbon markets are promising to finance coastal restoration, climate adaptation, and livelihoods for coastal communities.

Critical knowledge gaps need to be overcome before the full benefit of blue carbon can be realised. Priorities include: establishing the full global extent of BCEs and developing ongoing monitoring at management-relevant resolutions; addressing the permanence and temporal continuity of blue carbon storage and sequestration subject to SLR, changing climatic conditions and their impact on the distribution of mangroves, tidal wetlands, seagrass and macroalgae at high latitudes; establishing the factors that determine the carbon storage and sequestration capacity at the site scale and how these might be managed to increase mitigation benefit; and establishing the carbon mitigation potential and pathways for other coastal and marine ecosystems such as macroalgae and tidal forests. Given the recent momentum in blue carbon research, scientists and policy makers are well placed to address these gaps, providing research is sufficiently supported. Crucial to the effectiveness of blue carbon research for policy and management application is actively focussing on the highly under-studied regions, particularly in the global south, where the distribution of mangrove forests is greatest.

Open peer review

To view the open peer review materials for this article, please visit http://doi.org/10.1017/cft.2023.17.

Supplementary material

The supplementary material for this article can be found at http://doi.org/10.1017/cft.2023.17.

Data availability statement

Data were not generated to prepare this literature review.

Acknowledgements

The authors wish to acknowledge the traditional custodians of blue carbon estate globally and recognise their contribution to management and stewardship for millennia. We advocate for their leadership as blue carbon managers. Many collaborators have contributed to the research undertaken by the authors, and we appreciate their knowledge sharing.

Author contribution

Conception, writing and preparation of figures were principally undertaken by Rogers with substantial input for Kelleway and Saintilan. Figure 1 was prepared by Kelleway and Figures 4 and 5 by Rogers.

Financial support

We appreciate the financial support provided by respective institutions (University of Wollongong and Macquarie University), the Australian Research Council, and federal, state, and local governments who have supported the work of the authors.

References

Adame, MF, Cherian, S, Reef, R and Stewart-Koster, B (2017) Mangrove root biomass and the uncertainty of belowground carbon estimations. Forest Ecology and Management 403, 5260.CrossRefGoogle Scholar
Adame, M, Santini, N, Torres-Talamante, O and Rogers, K (2021) Mangrove sinkholes (cenotes) of the Yucatan peninsula, a global hotspot of carbon sequestration. Biology Letters 17(5), 20210037.CrossRefGoogle ScholarPubMed
Ahmed, N and Glaser, M (2016) Coastal aquaculture, mangrove deforestation and blue carbon emissions: Is REDD+ a solution? Marine Policy 66, 5866.CrossRefGoogle Scholar
Allen, JRL (2000) Morphodynamics of Holocene salt marshes: A review sketch from the Atlantic and southern North Sea coasts of Europe. Quaternary Science Reviews 19(12), 11551231.CrossRefGoogle Scholar
Alongi, DM (2012) Carbon sequestration in mangrove forests. Carbon Management 3(3), 313322.CrossRefGoogle Scholar
Alongi, DM (2020a) Carbon balance in salt marsh and mangrove ecosystems: A global synthesis. Journal of Marine Science and Engineering 8(10), 767.CrossRefGoogle Scholar
Alongi, DM (2020b) Global significance of mangrove blue carbon in climate change mitigation. Science 2(3), 67.CrossRefGoogle Scholar
Atwood, TB, Connolly, RM, Almahasheer, H, Carnell, PE, Duarte, CM, Ewers Lewis, CJ, Irigoien, X, Kelleway, JJ, Lavery, PS, Macreadie, PI, Serrano, O, Sanders, CJ, Santos, I, Steven, ADL and Lovelock, CE (2017) Global patterns in mangrove soil carbon stocks and losses. Nature Climate Change 7(7), 523528.CrossRefGoogle Scholar
Ball, MC (1988) Salinity tolerance in the mangroves Aegiceras corniculatum and Avicennia marina. I. Water use in relation to growth, carbon partitioning, and salt balance. Functional Plant Biology 15(3), 447464.CrossRefGoogle Scholar
Ball, MC (2002) Interactive effects of salinity and irradiance on growth: Implications for mangrove forest structure along salinity gradients. Trees 16(2), 126139.CrossRefGoogle Scholar
Ball, MC, Cochrane, MJ and Rawson, HM (1997) Growth and water use of the mangroves Rhizophora apiculata and R. stylosa in response to salinity and humidity under ambient and elevated concentrations of atmospheric CO2. Plant, Cell and Environment 20, 11581166.CrossRefGoogle Scholar
Beck, J, Bock, M, Schmitt, J, Seth, B, Blunier, T and Fischer, H (2018) Bipolar carbon and hydrogen isotope constraints on the Holocene methane budget. Biogeosciences 15(23), 71557175.CrossRefGoogle Scholar
Bertness, MD, Gough, L and Shumway, SW (1992) Salt tolerances and the distribution of fugitive salt marsh plants. Ecology 73(5), 18421851.CrossRefGoogle Scholar
Billah, MM, Bhuiyan, MKA, Islam, MA, Das, J and Hoque, ATMR (2022) Salt marsh restoration: An overview of techniques and success indicators. Environmental Science and Pollution Research 29(11), 15347–1563.CrossRefGoogle ScholarPubMed
Bindoff, NL, Cheung, WW, Kairo, JG, Arístegui, J, Guinder, VA, Hallberg, R, et al. (2019) Changing ocean, marine ecosystems, and dependent communities. In Pörtner, H-O, Roberts, DC, Masson-Delmotte, V, et al. (eds.), IPCC Special Report on the Ocean and Cryosphere in a Changing Climate. Cambridge, UK and New York, NY, USA: Cambridge University Press, pp. 447587.Google Scholar
Björk, M, Uku, J, Weil, A and Beer, S (1999) Photosynthetic tolerances to desiccation of tropical intertidal seagrasses. Marine Ecology Progress Series 191, 121126.CrossRefGoogle Scholar
Boski, T, Bezerra, FH, de Fátima, Pereira L, Souza, AM, Maia, RP and Lima-Filho, FP (2015) Sea-level rise since 8.2 ka recorded in the sediments of the Potengi–Jundiai Estuary, NE Brasil. Marine Geology 365, 113.CrossRefGoogle Scholar
Breithaupt, JL, Smoak, JM, Bianchi, TS, Vaughn, DR, Sanders, CJ, Radabaugh, KR, Osland, MJ, Feher, LC, Lynch, JC and Cahoon, DR (2020) Increasing rates of carbon burial in Southwest Florida coastal wetlands. Journal of Geophysical Research: Biogeosciences 125(2), e2019JG005349.Google Scholar
Bunting, P, Rosenqvist, A, Lucas, R, Rebelo, L-M, Hilarides, L, Thomas, N, Hardy, A, Itoh, T, Shimada, M and Finlayson, C (2018) The global mangrove watch—A new 2010 global baseline of mangrove extent. Remote Sensing 10(10), 1669.CrossRefGoogle Scholar
Cahoon, DR, McKee, KL and Morris, JT (2021) How plants influence resilience of salt marsh and mangrove wetlands to sea-level rise. Estuaries and Coasts 44(4), 883898.CrossRefGoogle Scholar
Campbell, AD, Fatoyinbo, L, Goldberg, L and Lagomasino, D (2022) Global hotspots of salt marsh change and carbon emissions. Nature 612, 701706.CrossRefGoogle ScholarPubMed
Canuel, EA and Hardison, AK (2016) Sources, ages, and alteration of organic matter in estuaries. Annual Review of Marine Science 8, 409434.CrossRefGoogle ScholarPubMed
Chmura, GL, Anisfeld, SC, Cahoon, DR and Lynch, JC (2003) Global carbon sequestration in tidal, saline wetland soils. Global Biogeochemical Cycles 17(4), 1111.CrossRefGoogle Scholar
Clark, JA, Farrell, WE and Peltier, WR (1978) Global changes in postglacial sea level: A numerical calculation. Quaternary Research 9(3), 265287.CrossRefGoogle Scholar
Cohen, KM, Cartelle, V, Barnett, R, Busschers, FS and Barlow, NL (2022) Last Interglacial Sea-level data points from Northwest Europe. Earth System Science Data 14(6), 28952937.CrossRefGoogle Scholar
Cohen, M, Filho, P, Lara, R, Behling, H and Angulo, R (2005) A model of Holocene mangrove development and Relative Sea-level changes on the Bragança Peninsula (Northern Brazil). Wetlands Ecology and Management 13(4), 433443.CrossRefGoogle Scholar
Collins, DS, Avdis, A, Allison, PA, Johnson, HD, Hill, J, Piggott, MD, Hassan, MHA and Damit, AR (2017) Tidal dynamics and mangrove carbon sequestration during the Oligo–Miocene in the South China Sea. Nature Communications 8(1), 112.CrossRefGoogle ScholarPubMed
Compton, JS (2001) Holocene Sea-level fluctuations inferred from the evolution of depositional environments of the southern Langebaan lagoon salt marsh, South Africa. The Holocene 11(4), 395405.CrossRefGoogle Scholar
Cooley, S, Schoeman, D, Bopp, L, Boyd, P, Donner, S, Ghebrehiwet, DY, Ito, S-I, Kiessling, W, Martinetto, P, Ojea, E, Racault, M-F, Rost, B, and Skern-Mauritzen, M (2022) Oceans and coastal ecosystems and their services. In Pörtner, H-O, Roberts, DC, Tignor, M, Poloczanska, ES, Mintenbeck, K, Alegría, A, Craig, M, Langsdorf, S, Löschke, S, Möller, V, Okem, A and Rama, B (eds.), Climate Change 2022: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, UK and New York, NY, USA: Cambridge University Press, pp. 379550. https://doi.org/10.1017/9781009325844.005Google Scholar
Creel, RC, Austermann, J, Khan, NS, D’Andrea, WJ, Balascio, N, Dyer, B, Ashe, E and Menke, W (2022) Postglacial relative sea level change in Norway. Quaternary Science Reviews 282, 107422.CrossRefGoogle Scholar
Darby, FA and Turner, RE (2008) Effects of eutrophication on salt marsh root and rhizome biomass accumulation. Marine Ecology Progress Series 363, 6370.CrossRefGoogle Scholar
DeLaune, R and White, J (2012) Will coastal wetlands continue to sequester carbon in response to an increase in global sea level?: A case study of the rapidly subsiding Mississippi river deltaic plain. Climatic Change 110(1), 297314.CrossRefGoogle Scholar
Digerfeldt, G and Hendry, M (1987) An 8000 year Holocene sea-level record from Jamaica: Implications for interpretation of Caribbean reef and coastal history. Coral Reefs 5(4), 165169.CrossRefGoogle Scholar
Donato, DC, Kauffman, JB, Murdiyarso, D, Kurnianto, S, Stidham, M and Kanninen, M (2011) Mangroves among the most carbon-rich forests in the tropics. Nature Geoscience 4(5), 293297.CrossRefGoogle Scholar
Duarte, CM and Cebrian, J (1996) The fate of marine autotrophic production. Limnology and Oceanography 41(8), 17581766.CrossRefGoogle Scholar
Duarte, CM, Losada, IJ, Hendriks, IE, Mazarrasa, I and Marba, N (2013) The role of coastal plant communities for climate change mitigation and adaptation. Nature Climate Change 3(11), 961968.CrossRefGoogle Scholar
Dynarski, KA, Bossio, DA and Scow, KM (2020) Dynamic stability of soil carbon: Reassessing the “permanence” of soil carbon sequestration. Frontiers in Environmental Science 8, 514701.CrossRefGoogle Scholar
Elsner, J, Jagger, T and Tsonis, A (2006) Estimated return periods for Hurricane Katrina. Geophysical Research Letters 33(8), L08704. https://doi.org/10.1029/2005GL025452.CrossRefGoogle Scholar
Emmer, I, Needelman, B, Emmett-Mattox, S, Crooks, S, Megonigal, P, Myers, D, Oreska, M, McGlathery, K and Shoch, DJVCS, Washington, DC (2015a) VM0033 Methodology for Tidal Wetland and Seagrass Restoration, v1.0.Google Scholar
Emmer, I, von Unger, M, Needelman, B, Crooks, S and Emmett-Mattox, S (2015b) Coastal Blue Carbon in Practice: A Manual for Using the VCS Methodology for Tidal Wetland and Seagrass Restoration VM0033. Arlington, Virginia: Restore America’s Estuaries and Silvestrum.Google Scholar
Fornari, M, Giannini, PCF and Junior, DRN (2012) Facies associations and controls on the evolution from a coastal bay to a lagoon system, Santa Catarina Coast, Brazil. Marine Geology 323, 5668.CrossRefGoogle Scholar
Forster, P, Storelvmo, T, Armour, K, Collins, W, Dufresne, J-L, Frame, D, Lunt, DJ, Mauritsen, T, Palmer, MD, Watanabe, M, Wild, M and Zhang, H (2021) The earth’s energy budget, climate feedbacks, and climate sensitivity. In Masson-Delmotte, V, Zhai, P, Pirani, A, Connors, SL, Péan, C, Berger, S, Caud, N, Chen, Y, Goldfarb, L, Gomis, MI, Huang, M, Leitzell, K, Lonnoy, E, Matthews, JBR, Maycock, TK, Waterfield, T, Yelekçi, O, Yu, R and Zhou, B (eds.), Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, UK and New York, NY, USA: Cambridge University Press, pp. 9231054. https://doi.org/10.1017/9781009157896.009Google Scholar
França, MC, Cohen, MC, Pessenda, LC, Rossetti, DF, Lorente, FL, Junior, AÁB, Guimarães, JT, Friaes, Y and Macario, K (2013) Mangrove vegetation changes on Holocene terraces of the Doce River, Southeastern Brazil. Catena 110, 5969.CrossRefGoogle Scholar
Friess, DA, Rogers, K, Lovelock, CE, Krauss, KW, Hamilton, SE, Lee, SY, Lucas, R, Primavera, J, Rajkaran, A and Shi, S (2019) The state of the World’s Mangrove Forests: Past, present, and future. Annual Review of Environment and Resources 44(1), 89115.CrossRefGoogle Scholar
Friess, DA, Yando, ES, Abuchahla, GM, Adams, JB, Cannicci, S, Canty, SW, Cavanaugh, KC, Connolly, RM, Cormier, N and Dahdouh-Guebas, F (2020) Mangroves give cause for conservation optimism, for now. Current Biology 30(4), R153R154.CrossRefGoogle ScholarPubMed
Galik, C, Baker, JS, Daigneault, A and Latta, G (2022) Crediting temporary forest carbon: Retrospective and empirical perspectives on accounting options. Frontiers in Forests and Global Change Biology 5, 933020.CrossRefGoogle Scholar
Gedan, KB, Silliman, BR and Bertness, MD (2009) Centuries of human-driven change in salt marsh ecosystems. Annual Review of Marine Science 1(1), 117141.CrossRefGoogle Scholar
Gehrels, WR (1999) Middle and late Holocene Sea-level changes in eastern Maine reconstructed from foraminiferal saltmarsh stratigraphy and AMS 14C dates on basal peat. Quaternary Research 52(3), 350359.CrossRefGoogle Scholar
Godoy, MD and de Lacerda, LD (2015) Mangroves response to climate change: A review of recent findings on mangrove extension and distribution. Anais da Academia Brasileira de Ciências 87, 651667.CrossRefGoogle ScholarPubMed
Goldberg, L, Lagomasino, D, Thomas, N and Fatoyinbo, T (2020) Global declines in human‐driven mangrove loss. Global Change Biology 26(10), 58445855.CrossRefGoogle Scholar
Greiner, JT, McGlathery, KJ, Gunnell, J and McKee, BA (2013) Seagrass restoration enhances “blue carbon” sequestration in coastal waters. PLoS One 8(8), e72469.CrossRefGoogle ScholarPubMed
Grindrod, J (2002) Late quaternary mangrove pollen records from continental shelf and ocean cores in the north Australian-Indonesian region. In Kershaw, P, David, B, Tapper, N, Penny, D and Brown, J (eds.), Bridging Wallace’s Line: The Environmental and Cultural History and Dynamics of the Australian-Southeast Asian Region. Reiskirchen Germany: Catena-Verlag, pp. 119146.Google Scholar
Grindrod, J, Moss, P and Kaars, SVD (1999) Late quaternary cycles of mangrove development and decline on the north Australian continental shelf. Journal of Quaternary Science 14(5), 465470.3.0.CO;2-E>CrossRefGoogle Scholar
Griscom, BW, Adams, J, Ellis, PW, Houghton, RA, Lomax, G, Miteva, DA, Schlesinger, WH, Shoch, D, Siikamäki, JV, Smith, P, Woodbury, P, Zganjar, C, Blackman, A, Campari, J, Conant, RT, Delgado, C, Elias, P, Gopalakrishna, T, Hamsik, MR, Herrero, M, Kiesecker, J, Landis, E, Laestadius, L, Leavitt, SM, Minnemeyer, S, Polasky, S, Potapov, P, Putz, FE, Sanderman, J, Silvius, M, Wollenberg, E and Fargione, J (2017) Natural climate solutions. Proceedings of the National Academy of Sciences 114(44), 11645.CrossRefGoogle ScholarPubMed
Hagger, V, Waltham, NJ and Lovelock, CE (2022) Opportunities for coastal wetland restoration for blue carbon with co-benefits for biodiversity, coastal fisheries, and water quality. Ecosystem Services 55, 101423.CrossRefGoogle Scholar
Haigh, ID, Dornbusch, U, Brown, J, Lyddon, C, Nicholls, RJ, Penning-Roswell, E and Sayers, P (2022) Climate change impacts on coastal flooding relevant to the UK and Ireland. MCCIP Science Review 2022, 118.Google Scholar
Hait, AK and Behling, H (2009) Holocene mangrove and coastal environmental changes in the Western Ganga–Brahmaputra Delta, India. Vegetation History and Archaeobotany 18(2), 159169.CrossRefGoogle Scholar
Hanebuth, T, Stattegger, K and Grootes, PM (2000) Rapid flooding of the Sunda Shelf: A late-glacial sea-level record. Science 288(5468), 10331035.CrossRefGoogle Scholar
Hashimoto, TR, Saintilan, N and Haberle, SG (2006) Mid-Holocene development of mangrove communities featuring Rhizophoraceae and geomorphic change in the Richmond River Estuary, New South Wales, Australia. Geographical Research 44(1), 6376.CrossRefGoogle Scholar
Haywood, BJ, Hayes, MP, White, JR and Cook, RL (2020) Potential fate of wetland soil carbon in a deltaic coastal wetland subjected to high relative sea level rise. Science of the Total Environment 711, 135185.CrossRefGoogle Scholar
Herr, D, Blum, J, Himes-Cornell, A and Sutton-Grier, A (2019) An analysis of the potential positive and negative livelihood impacts of coastal carbon offset projects. Journal of Environmental Management 235, 463479.CrossRefGoogle ScholarPubMed
Hopkinson, CS, Morris, JT, Fagherazzi, S, Wollheim, WM and Raymond, PA (2018) Lateral marsh edge erosion as a source of sediments for vertical marsh accretion. Journal of Geophysical Research: Biogeosciences 123(8), 24442465.CrossRefGoogle Scholar
Horton, BP, Gibbard, PL, Mine, G, Morley, R, Purintavaragul, C and Stargardt, JM (2005) Holocene sea levels and palaeoenvironments, Malay-Thai Peninsula, Southeast Asia. The Holcene 15(8), 11991213.CrossRefGoogle Scholar
Horton, BP, Shennan, I, Bradley, SL, Cahill, N, Kirwan, M, Kopp, RE and Shaw, TA (2018) Predicting marsh vulnerability to sea-level rise using Holocene relative sea-level data. Nature Communications 9(1), 17.CrossRefGoogle ScholarPubMed
Howard, JL, Creed, JC, Aguiar, MV and Fourqurean, JW (2018) CO2 released by carbonate sediment production in some coastal areas may offset the benefits of seagrass “blue carbon” storage. Limnology and Oceanography 63(1), 160172.CrossRefGoogle Scholar
Howard, J, Hoyt, S, Isensee, K, Pidgeon, E and Telszewski, M (2014) Coastal Blue Carbon: Methods for Assessing Carbon Stocks and Emissions Factors in Mangroves, Tidal Salt Marshes, and Seagrass Meadows. Arlington, Virginia, USA: Conservation International, Intergovernmental Oceanographic Commission of UNESCO, International Union for Conservation of Nature.Google Scholar
IPCC (2021) Summary for policymakers. Climate change 2021: The physical science basis. In Masson-Delmotte, V, Zhai, P, Pirani, A, et al. (eds), Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge: Cambridge University Press.Google Scholar
Iyer, GC, Edmonds, JA, Fawcett, AA, Hultman, NE, Alsalam, J, Asrar, GR, Calvin, KV, Clarke, LE, Creason, J and Jeong, M (2015) The contribution of Paris to limit global warming to 2°C. Environmental Research Letters 10(12), 125002.CrossRefGoogle Scholar
Jayathilake, DRM and Costello, MJ (2018) A modelled global distribution of the seagrass biome. Biological Conservation 226, 120126.CrossRefGoogle Scholar
Jeffrey, LC, Reithmaier, G, Sippo, JZ, Johnston, SG, Tait, DR, Harada, Y and Maher, DT (2019) Are methane emissions from mangrove stems a cryptic carbon loss pathway? Insights from a catastrophic forest mortality. New Phytologist 224(1), 146154.CrossRefGoogle ScholarPubMed
Jervey, M (1988) Quantitative geological modeling of siliciclastic rock sequences and their seismic expression. In Wilgus, CK, Hastings, BS, Posamentier, H, et al (eds.), Sea Level Changes: An Integrated Approach: Society of Economic Paleontologists and Mineralogists Special Publication, vol. 42. Tulsa, Oklahoma: SEPM (Society for Sedimentary Geology), pp. 4769.CrossRefGoogle Scholar
Jin-Eong, O, Khoon, GW and Clough, B (1995) Structure and productivity of a 20-year-old stand of Rhizophora apiculata Bl. Mangrove forest. Journal of Biogeography 22(2/3), 417424.CrossRefGoogle Scholar
Jones, MC, Wingard, GL, Stackhouse, B, Keller, K, Willard, D, Marot, M, Landacre, B and Bernhardt, CE (2019) Rapid inundation of southern Florida coastline despite low relative sea-level rise rates during the late-Holocene. Nature Communications 10(1), 3231.CrossRefGoogle ScholarPubMed
Kearney, MS, Stevenson, JC and Ward, LG (1994) Spatial and temporal changes in marsh vertical accretion rates at Monie Bay: Implications for sea-level rise. Journal of Coastal Research 10(4), 10101020.Google Scholar
Kelleway, JJ, Saintilan, N, Macreadie, PI and Ralph, PJ (2016a) Sedimentary factors are key predictors of carbon storage in SE Australian saltmarshes. Ecosystems 19(5), 865880.CrossRefGoogle Scholar
Kelleway, JJ, Saintilan, N, Macreadie, PI, Skilbeck, CG, Zawadzki, A and Ralph, PJ (2016b) Seventy years of continuous encroachment substantially increases ‘blue carbon’ capacity as mangroves replace intertidal salt marshes. Global Change Biology 22(3), 10971109.CrossRefGoogle ScholarPubMed
Kennedy, D, Wong, V and Jacobsen, G (2021) Holocene infill of the Anglesea Estuary, Victoria: A keep-up estuary in a geologically constrained environment. Australian Journal of Earth Sciences 68, 113.CrossRefGoogle Scholar
Kerswell, AP (2006) Global biodiversity patterns of benthic marine algae. Ecology 87(10), 24792488.CrossRefGoogle ScholarPubMed
Khan, NS, Ashe, E, Shaw, TA, Vacchi, M, Walker, J, Peltier, W, Kopp, RE and Horton, BP (2015) Holocene relative sea-level changes from near-, intermediate-, and far-field locations. Current Climate Change Reports 1(4), 247262.CrossRefGoogle Scholar
Kihara, Y, Dannoura, M and Ohashi, M (2022) Estimation of fine root production, mortality, and decomposition by using two core methods and litterbag experiments in a mangrove forest. Ecological Research 37(1), 5366.CrossRefGoogle Scholar
Kirwan, ML, Guntenspergen, GR and Langley, J (2014) Temperature sensitivity of organic-matter decay in tidal marshes. Biogeosciences 11(17), 48014808.CrossRefGoogle Scholar
Kirwan, ML and Mudd, SM (2012) Response of salt-marsh carbon accumulation to climate change. Nature 489(7417), 550553.CrossRefGoogle ScholarPubMed
Koch, EW (2001) Beyond light: Physical, geological, and geochemical parameters as possible submersed aquatic vegetation habitat requirements. Estuaries 24(1), 117.CrossRefGoogle Scholar
Komiyama, A, Ong, JE and Poungparn, S (2008) Allometry, biomass, and productivity of mangrove forests: A review. Aquatic Botany 89(2), 128137.CrossRefGoogle Scholar
Krause-Jensen, D and Duarte, CM (2016) Substantial role of macroalgae in marine carbon sequestration. Nature Geoscience 9(10), 737742.CrossRefGoogle Scholar
Krauss, KW, Demopoulos, AW, Cormier, N, From, A, McClain-Counts, JP and Lewis, RR (2018) Ghost forests of Marco Island: Mangrove mortality driven by belowground soil structural shifts during tidal hydrologic alteration. Estuarine, Coastal and Shelf Science 212, 5162.CrossRefGoogle Scholar
Kroeger, KD, Crooks, S, Moseman-Valtierra, S and Tang, J (2017) Restoring tides to reduce methane emissions in impounded wetlands: A new and potent blue carbon climate change intervention. Scientific Reports 7(1), 11914.CrossRefGoogle ScholarPubMed
Lambeck, K, Rouby, H, Purcell, A, Sun, Y and Sambridge, M (2014) Sea level and global ice volumes from the last glacial maximum to the Holocene. Proceedings of the National Academy of Sciences 111(43), 1529615303.CrossRefGoogle ScholarPubMed
Lamont, K, Saintilan, N, Kelleway, JJ, Mazumder, D and Zawadzki, A (2020) Thirty-year repeat measures of mangrove above-and below-ground biomass reveals unexpectedly high carbon sequestration. Ecosystems 23(2), 370382.CrossRefGoogle Scholar
Langley, JA, McKee, KL, Cahoon, DR, Cherry, JA and Megonigal, JP (2009) Elevated CO2 stimulates marsh elevation gain, counterbalancing sea-level rise. Proceedings of the National Academy of Sciences 106(15), 61826186.CrossRefGoogle Scholar
Lee, SY, Hamilton, S, Barbier, EB, Primavera, J and Lewis, RR (2019) Better restoration policies are needed to conserve mangrove ecosystems. Nature Ecology and Evolution 3(6), 870872.CrossRefGoogle ScholarPubMed
Li, Z, Saito, Y, Mao, L, Tamura, T, Song, B, Zhang, Y, Lu, A, Sieng, S and Li, J (2012) Mid-Holocene mangrove succession and its response to sea-level change in the upper Mekong River delta, Cambodia. Quaternary Research 78(2), 386399.CrossRefGoogle Scholar
Lichacz, W, Hardiman, S and Buckney, R (2009) Below-ground biomass in some intertidal wetlands in New South Wales. Wetlands(Australia) 4(2), 5662.Google Scholar
Lovelock, CE, Adame, MF, Bradley, J, Dittmann, S, Hagger, V, Hickey, SM, Hutley, LB, Jones, A, Kelleway, JJ and Lavery, PS (2022a) An Australian blue carbon method to estimate climate change mitigation benefits of coastal wetland restoration. Restoration Ecology, e13739. https://doi.org/10.1111/rec.13739CrossRefGoogle Scholar
Lovelock, CE, Adame, MF, Butler, DW, Kelleway, JJ, Dittmann, S, Fest, B, King, KJ, Macreadie, PI, Mitchell, K and Newnham, MJGEC (2022b) Modeled approaches to estimating blue carbon accumulation with mangrove restoration to support a blue carbon accounting method for Australia. Limnology and Oceanography 67, S50S60.CrossRefGoogle Scholar
Lovelock, CE, Barbier, E and Duarte, CM (2022c) Tackling the mangrove restoration challenge. PLoS Biology 20(10), e3001836.CrossRefGoogle ScholarPubMed
Lovelock, CE and Duarte, CM (2019) Dimensions of blue carbon and emerging perspectives. Biology Letters 15(3), 20180781.CrossRefGoogle ScholarPubMed
Lovelock, CE and Reef, R (2020) Variable impacts of climate change on blue carbon. One Earth 3(2), 195211.CrossRefGoogle Scholar
Ma, S, He, F, Tian, D, Zou, D, Yan, Z, Yang, Y, Zhou, T, Huang, K, Shen, H and Fang, J (2018) Variations and determinants of carbon content in plants: A global synthesis. Biogeosciences 15(3), 693702.CrossRefGoogle Scholar
Mack, SK, Lane, RR, Deng, J, Morris, JT and Bauer, JJ (2023) Wetland carbon models: Applications for wetland carbon commercialization. Ecological Modelling 476, 110228.CrossRefGoogle Scholar
Macreadie, PI, Anton, A, Raven, JA, Beaumont, N, Connolly, RM, Friess, DA, Kelleway, JJ, Kennedy, H, Kuwae, T, Lavery, PS, Lovelock, CE, Smale, DA, Apostolaki, ET, Atwood, TB, Baldock, J, Bianchi, TS, Chmura, GL, Eyre, BD, Fourqurean, JW, Hall-Spencer, JM, Huxham, M, Hendriks, IE, Krause-Jensen, D, Laffoley, D, Luisetti, T, Marbà, N, Masque, P, KJ, MG, Megonigal, JP, Murdiyarso, D, Russell, BD, Santos, R, Serrano, O, Silliman, BR, Watanabe, K and Duarte, CM (2019) The future of blue carbon science. Nature Communications 10(1), 3998.CrossRefGoogle ScholarPubMed
Macreadie, PI, Costa, MD, Atwood, TB, Friess, DA, Kelleway, JJ, Kennedy, H, Lovelock, CE, Serrano, O and Duarte, CM (2021) Blue carbon as a natural climate solution. Nature Reviews Earth and Environment 2(12), 826839.CrossRefGoogle Scholar
Macreadie, PI, Robertson, AI, Spinks, B, Adams, MP, Atchison, JM, Bell-James, J, Bryan, BA, Chu, L, Filbee-Dexter, K and Drake, L (2022) Operationalizing marketable blue carbon. One Earth 5(5), 485492.CrossRefGoogle Scholar
Madsen, JD, Chambers, PA, James, WF, Koch, EW and Westlake, DF (2001) The interaction between water movement, sediment dynamics and submersed macrophytes. Hydrobiologia 444(1), 7184.CrossRefGoogle Scholar
Maher, DT, Call, M, Santos, IR and Sanders, CJ (2018) Beyond burial: Lateral exchange is a significant atmospheric carbon sink in mangrove forests. Biology Letters 14(7), 20180200.CrossRefGoogle ScholarPubMed
Malerba, ME, Friess, DA, Peacock, M, Grinham, A, Taillardat, P, Rosentreter, JA, Webb, J, Iram, N, Al-Haj, AN and Macreadie, PI (2022) Methane and nitrous oxide emissions complicate the climate benefits of teal and blue carbon wetlands. One Earth 5(12), 13361341.CrossRefGoogle Scholar
Marx, SK, Knight, JM, Dwyer, PG, Child, DP, Hotchkis, MA and Zawadzki, A (2020) Examining the response of an eastern Australian mangrove forest to changes in hydro-period over the last century. Estuarine, Coastal and Shelf Science 241, 106813.CrossRefGoogle Scholar
McKee, KL (2011) Biophysical controls on accretion and elevation change in Caribbean mangrove ecosystems. Estuarine, Coastal and Shelf Science 91(4), 475483.CrossRefGoogle Scholar
McKee, KL, Cahoon, DR and Feller, IC (2007) Caribbean mangroves adjust to rising sea level through biotic controls on change in soil elevation. Global Ecology and Biogeography 16(5), 545556.CrossRefGoogle Scholar
McKee, K, Rogers, K and Saintilan, N (2012) Response of salt marsh and mangrove wetlands to changes in atmospheric CO2, climate, and sea level. In Middleton, BA (ed.), Global Change and the Function and Distribution of Wetlands, vol. 1. Netherlands: Springer, pp. 6396.CrossRefGoogle Scholar
McLeod, E, Chmura, GL, Bouillon, S, Salm, R, Björk, M, Duarte, CM, Lovelock, CE, Schlesinger, WH and Silliman, BR (2011) A blueprint for blue carbon: Toward an improved understanding of the role of vegetated coastal habitats in sequestering CO2. Frontiers in Ecology and the Environment 9(10), 552560.CrossRefGoogle Scholar
Mcowen, C, Weatherdon, L, Bochove, J-W, Sullivan, E, Blyth, S, Zockler, C, Stanwell-Smith, D, Kingston, N, Martin, C, Spalding, M and Fletcher, S (2017) A global map of saltmarshes. Biodiversity Data Journal 5, e11764.CrossRefGoogle Scholar
Michael, PS (2013) Ecological impacts and management of acid sulphate soil: A review. Asian Journal of Water, Environment and Pollution 10(4), 1324.Google Scholar
Morris, JT, Sundareshwar, PV, Nietch, CT, Kjerfve, B and Cahoon, DR (2002) Responses of coastal wetlands to rising sea-levels. Ecology 83(10), 28692877.CrossRefGoogle Scholar
Mudd, SM, Howell, SM and Morris, JT (2009) Impact of dynamic feedbacks between sedimentation, sea-level rise, and biomass production on near-surface marsh stratigraphy and carbon accumulation. Estuarine, Coastal and Shelf Science 82(3), 377389.CrossRefGoogle Scholar
Mueller, P, Schile-Beers, LM, Mozdzer, TJ, Chmura, GL, Dinter, T, Kuzyakov, Y, de Groot, AV, Esselink, P, Smit, C and D’Alpaos, A (2018) Global-change effects on early-stage decomposition processes in tidal wetlands–implications from a global survey using standardized litter. Biogeosciences 15(10), 31893202.CrossRefGoogle Scholar
Muhammad-Nor, SM, Huxham, M, Salmon, Y, Duddy, SJ, Mazars-Simon, A, Mencuccini, M, Meir, P and Jackson, G (2019) Exceptionally high mangrove root production rates in the Kelantan Delta, Malaysia; an experimental and comparative study. Forest Ecology and Management 444, 214224.CrossRefGoogle Scholar
Murray, NJ, Worthington, TA, Bunting, P, Duce, S, Hagger, V, Lovelock, CE, Lucas, R, Saunders, MI, Sheaves, M and Spalding, MJS (2022) High-resolution mapping of losses and gains of Earth’s tidal wetlands. Science 376(6594), 744749.CrossRefGoogle ScholarPubMed
Murray-Wallace, CV and Woodroffe, CD (2014) Quaternary Sea-Level Changes: A Global Perspective. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Neumann, AC and Macintyre, IG (1985) Reef response of sea level rise: Keep-up, catch-up or give-up. In Gabrie, C, Toffart, JL and Salvat, B (eds.), Proceedings of the Fifth International Coral Reef Congress, Tahiti, 27 May–1 June 1985: Volume 3, Symposia and Seminars (A), pp. 105110.Google Scholar
Nicholas, W, Nichol, S, Howard, F, Picard, K, Dulfer, H, Radke, L, Carroll, A, Tran, M and Siwabessy, P (2014) Pockmark development in the petrel sub-basin, Timor Sea, Northern Australia: Seabed habitat mapping in support of CO2 storage assessments. Continental Shelf Research 83, 129142.CrossRefGoogle Scholar
Orson, RA, Warren, RS and Niering, WA (1998) Interpreting sea level rise and rates of vertical marsh accretion in a Southern New England tidal salt marsh. Estuarine, Coastal and Shelf Science 47(4), 419429.CrossRefGoogle Scholar
Osland, MJ, Enwright, NM, Day, RH, Gabler, CA, Stagg, CL and Grace, JB (2016) Beyond just sea‐level rise: Considering macroclimatic drivers within coastal wetland vulnerability assessments to climate change. Global Change Biology 22(1), 111.CrossRefGoogle ScholarPubMed
Osland, MJ, Feher, LC, Griffith, KT, Cavanaugh, KC, Enwright, NM, Day, RH, Stagg, CL, Krauss, KW, Howard, RJ, Grace, JB and Rogers, K (2017) Climatic controls on the global distribution, abundance, and species richness of mangrove forests. Ecological Monographs 87(2), 341359.CrossRefGoogle Scholar
Osland, MJ, Feher, LC, Spivak, AC, Nestlerode, JA, Almario, AE, Cormier, N, From, AS, Krauss, KW, Russell, MJ and Alvarez, F (2020) Rapid peat development beneath created, maturing mangrove forests: Ecosystem changes across a 25‐yr chronosequence. Ecological Applications 30(4), e02085.CrossRefGoogle Scholar
Ouyang, X and Lee, S (2014) Updated estimates of carbon accumulation rates in coastal marsh sediments. Biogeosciences 11(18), 50575071.CrossRefGoogle Scholar
Owers, CJ, Woodroffe, CD, Mazumder, D and Rogers, K (2022) Carbon storage in coastal wetlands is related to elevation and how it changes over time. Estuarine, Coastal and Shelf Science 267, 107775.CrossRefGoogle Scholar
Parkinson, RW (1989) Decelerating holocene sea-level rise and its influence on Southwest Florida coastal evolution; A transgressive/regressive stratigraphy. Journal of Sedimentary Research 59(6), 960972.Google Scholar
Parkinson, RW, Ron, DD and White, JR (1994) Holocene sea-level rise and the fate of mangrove forests within the wider Caribbean Region. Journal of Coastal Research 10(4), 10771086.Google Scholar
Pendleton, L, Donato, DC, Murray, BC, Crooks, S, Jenkins, WA, Sifleet, S, Craft, C, Fourqurean, JW, Kauffman, JB and Marbà, N (2012) Estimating global “blue carbon” emissions from conversion and degradation of vegetated coastal ecosystems. PLoS One 7(9), e43542.CrossRefGoogle ScholarPubMed
Pham, TD, Xia, J, Ha, NT, Bui, DT, Le, NN and Tekeuchi, W (2019) A review of remote sensing approaches for monitoring blue carbon ecosystems: Mangroves, Seagrassesand salt marshes during 2010–2018. Sensors 19(8), 1933.CrossRefGoogle ScholarPubMed
Poffenbarger, H, Needelman, B and Megonigal, J (2011) Salinity influence on methane emissions from tidal marshes. Wetlands 31(5), 831842.CrossRefGoogle Scholar
Pons, L, Van Breemen, N and Driessen, P (1982) Physiography of coastal sediments and development of potential soil acidity. Acid Sulfate Weathering 10, 118.Google Scholar
Proske, U and Haberle, SG (2012) Island ecosystem and biodiversity dynamics in Northeastern Australia during the Holocene: Unravelling short-term impacts and long-term drivers. The Holocene 22(10), 10971111.CrossRefGoogle Scholar
Punwong, P, Selby, K and Marchant, R (2018) Holocene mangrove dynamics and relative sea-level changes along the Tanzanian Coast, East Africa. Estuarine, Coastal and Shelf Science 212, 105117.CrossRefGoogle Scholar
Quisthoudt, K, Schmitz, N, Randin, C, Dahdouh-Guebas, F, Robert, ER and Koedam, N (2012) Temperature variation among mangrove latitudinal range limits worldwide. Trees 26(6), 19191931.CrossRefGoogle Scholar
Radabaugh, KR, Powell, CE, Bociu, I, Clark, BC and Moyer, RP (2017) Plant size metrics and organic carbon content of Florida salt marsh vegetation. Wetlands Ecology and Management 25(4), 443455.CrossRefGoogle Scholar
Rahman, MM, Zimmer, M, Ahmed, I, Donato, D, Kanzaki, M and Xu, M (2021) Co-benefits of protecting mangroves for biodiversity conservation and carbon storage. Nature Communications 12(1), 19.CrossRefGoogle ScholarPubMed
Raven, J (2018) Blue carbon: Past, present and future, with emphasis on macroalgae. Biology Letters 14(10), 20180336.CrossRefGoogle ScholarPubMed
Raw, J, Riddin, T, Wasserman, J, Lehman, T, Bornman, T and Adams, J (2020) Salt marsh elevation and responses to future sea-level rise in the Knysna Estuary, South Africa. African Journal of Aquatic Science 45(1–2), 4964.CrossRefGoogle Scholar
Redfield, AC (1972) Development of a New England salt marsh. Ecological Monographs 42(2), 201237.CrossRefGoogle Scholar
Reed, DJ (1990) The impact of sea-level rise on coastal salt marshes. Progress in Physical Geography 14(4), 465481.CrossRefGoogle Scholar
Reef, R, Spencer, T, Mӧller, I, Lovelock, CE, Christie, EK, McIvor, AL, Evans, BR and Tempest, JA (2017) The effects of elevated CO2 and eutrophication on surface elevation gain in a European salt marsh. Global Change Biology 23(2), 881890.CrossRefGoogle Scholar
Ribeiro, SR, Batista, E.J.L, Cohen, M.C.L, França, MC, Pessenda, LCR, Fontes, NA, Alves, ICC and Bendassolli, JA (2018) Allogenic and autogenic effects on mangrove dynamics from the Ceará Mirim River, North-Eastern Brazil, during the middle and late Holocene. Earth Surface Processes and Landforms 43(8), 16221635.CrossRefGoogle Scholar
Rogers, K (2021) Accommodation space as a framework for assessing the response of mangroves to relative sea-level rise. Singapore Journal of Tropical Geography 42(2), 163183.CrossRefGoogle Scholar
Rogers, K, Kelleway, JJ, Saintilan, N, Megonigal, JP, Adams, JB, Holmquist, JR, Lu, M, Schile-Beers, L, Zawadzki, A, Mazumder, D and Woodroffe, CD (2019a) Wetland carbon storage controlled by millennialscale variation in relative sea-level rise. Nature 567, 9195.CrossRefGoogle ScholarPubMed
Rogers, K, Lal, KK, Asbridge, EF and Dwyer, PG (2022) Coastal wetland rehabilitation first-pass prioritisation for blue carbon and associated co-benefits. Marine and Freshwater Research. https://doi.org/10.1071/MF22014CrossRefGoogle Scholar
Rogers, K, Macreadie, PI, Kelleway, JJ and Saintilan, N (2019b) Blue carbon in coastal landscapes: A spatial framework for assessment of stocks and additionality. Sustainability Science 14(2), 453467.CrossRefGoogle Scholar
Rogers, K and Woodroffe, CD (2014) Tidal flats and salt marshes. In Masselink, G and Gehrels, R (eds.), Coastal Environments and Global Change. Oxford, UK: Wiley and Sons.Google Scholar
Romañach, SS, DeAngelis, DL, Koh, HL, Li, Y, Teh, SY, Barizan, RSR and Zhai, L (2018) Conservation and restoration of mangroves: Global status, perspectives, and prognosis. Ocean and Coastal Management 154, 7282.CrossRefGoogle Scholar
Rosentreter, JA, Al‐Haj, AN, Fulweiler, RW and Williamson, P (2021a) Methane and nitrous oxide emissions complicate coastal blue carbon assessments. Global Biogeochemical Cycles 35(2), e2020GB006858.CrossRefGoogle Scholar
Rosentreter, JA, Borges, AV, Deemer, BR, Holgerson, MA, Liu, S, Song, C, Melack, J, Raymond, PA, Duarte, CM and Allen, GH (2021b) Half of global methane emissions come from highly variable aquatic ecosystem sources. Nature Geoscience 14(4), 225230.CrossRefGoogle Scholar
Rosentreter, JA, Maher, DT, Erler, DV, Murray, RH and Eyre, BD (2018) Methane emissions partially offset “blue carbon” burial in mangroves. Science Advances 4(6), eaao4985.CrossRefGoogle ScholarPubMed
Rovai, A, Riul, P, Twilley, R, Castañeda‐Moya, E, Rivera‐Monroy, V, Williams, A, Simard, M, Cifuentes‐Jara, M, Lewis, R and Crooks, S (2016) Scaling mangrove aboveground biomass from site‐level to continental‐scale. Global Ecology and Biogeography 25(3), 286298.CrossRefGoogle Scholar
Rovai, AS, Twilley, RR, Castañeda-Moya, E, Riul, P, Cifuentes-Jara, M, Manrow-Villalobos, M, Horta, PA, Simonassi, JC, Fonseca, AL and Pagliosa, PR (2018) Global controls on carbon storage in mangrove soils. Nature Climate Change 8, 534538.CrossRefGoogle Scholar
Sadat-Noori, M, Rankin, C, Rayner, D, Heimhuber, V, Gaston, T, Drummond, C, Chalmers, A, Khojasteh, D and Glamore, W (2021) Coastal wetlands can be saved from sea level rise by recreating past tidal regimes. Scientific Reports 11(1), 1196.CrossRefGoogle ScholarPubMed
Saderne, V, Geraldi, NR, Macreadie, PI, Maher, DT, Middelburg, J, Serrano, O, Almahasheer, H, Arias-Ortiz, A, Cusack, M and BDJNc, E (2019) Role of carbonate burial in Blue Carbon budgets. Nature Communications 10(1), 19.CrossRefGoogle ScholarPubMed
Saintilan, N (1997) Above- and below-ground biomass of mangroves in a sub-tropical estuary. Marine and Freshwater Research 48(7), 601604.CrossRefGoogle Scholar
Saintilan, N, Khan, N, Ashe, E, Kelleway, J, Rogers, K, Woodroffe, CD and Horton, B (2020) Thresholds of mangrove survival under rapid sea level rise. Science 368(6495), 11181121.CrossRefGoogle ScholarPubMed
Saintilan, N, Kovalenko, KE, Guntenspergen, G, Rogers, K, Lynch, JC, Cahoon, DR, Lovelock, CE, Friess, DA, Ashe, E and Krauss, KW (2022) Constraints on the adjustment of tidal marshes to accelerating sea level rise. Science 377(6605), 523527.CrossRefGoogle ScholarPubMed
Saintilan, N, Rogers, K, Mazumder, D and Woodroffe, C (2013) Allochthonous and autochthonous contributions to carbon accumulation and carbon store in Southeastern Australian coastal wetlands. Estuarine, Coastal and Shelf Science 128, 8492.CrossRefGoogle Scholar
Saintilan, N, Wilson, NC, Rogers, K, Rajkaran, A and Krauss, KW (2014) Mangrove expansion and salt marsh decline at mangrove poleward limits. Global Change Biology 20, 147157.CrossRefGoogle ScholarPubMed
Sanderman, J, Hengl, T, Fiske, G, Solvik, K, Adame, MF, Benson, L, Bukoski, JJ, Carnell, P, Cifuentes-Jara, M and Donato, D (2018) A global map of mangrove forest soil carbon at 30 m spatial resolution. Environmental Research Letters 13(5), 055002.CrossRefGoogle Scholar
Sanders, CJ, Maher, DT, Tait, DR, Williams, D, Holloway, C, Sippo, JZ and Santos, IR (2016) Are global mangrove carbon stocks driven by rainfall? Journal of Geophysical Research: Biogeosciences 121(10), 26002609.CrossRefGoogle Scholar
Sani, DA, Hashim, M and Hossain, MS (2019) Recent advancement on estimation of blue carbon biomass using satellite-based approach. International Journal of Remote Sensing 40(20), 76797715.CrossRefGoogle Scholar
Sapkota, Y and White, JR (2020) Carbon offset market methodologies applicable for coastal wetland restoration and conservation in the United States: A review. Science of the Total Environment 701, 134497.CrossRefGoogle ScholarPubMed
Scholl, DW (1964) Recent sedimentary record in mangrove swamps and rise in sea level over the Southwestern Coast of Florida: Part 1. Marine Geology 1(4), 344366.CrossRefGoogle Scholar
Schuerch, M, Spencer, T, Temmerman, S, Kirwan, ML, Wolff, C, Lincke, D, McOwen, CJ, Pickering, MD, Reef, R, Vafeidis, AT, Hinkel, J, Nicholls, RJ and Brown, S (2018) Future response of global coastal wetlands to sea-level rise. Nature 561(7722), 231234.CrossRefGoogle ScholarPubMed
Sefton, JP, Woodroffe, SA and Ascough, P (2021) Radiocarbon dating of mangrove sediments. In Friess, DA and Sidik, F (eds) Dynamic Sedimentary Environments of Mangrove Coasts. Amsterdam: Elsevier, pp. 199210.CrossRefGoogle Scholar
Sherrod, CL and McMillan, C (1985) The distributional history and ecology of mangrove vegetation along the northern Gulf of Mexico coastal region. Contributions in Marine Science 28, 129140Google Scholar
Silvestri, S, Defina, A and Marani, M (2005) Tidal regime, salinity and salt marsh plant zonation. Estuarine, Coastal and Shelf Science 62(1–2), 119130.CrossRefGoogle Scholar
Simard, M, Fatoyinbo, L, Smetanka, C, Rivera-Monroy, VH, Castañeda-Moya, E, Thomas, N and Van der Stocken, T (2019) Mangrove canopy height globally related to precipitation, temperature and cyclone frequency. Nature Geoscience 12(1), 40.CrossRefGoogle Scholar
Simpson, L, Stein, C, Osborne, T and Feller, I (2019) Mangroves dramatically increase carbon storage after 3 years of encroachment. Hydrobiologia 834(1), 1326.CrossRefGoogle Scholar
Sowers, T (2010) Atmospheric methane isotope records covering the Holocene period. Quaternary Science Reviews 29(1–2), 213221.CrossRefGoogle Scholar
Spivak, AC, Sanderman, J, Bowen, JL, Canuel, EA and Hopkinson, CS (2019) Global-change controls on soil-carbon accumulation and loss in coastal vegetated ecosystems. Nature Geoscience 12(9), 685692.CrossRefGoogle Scholar
Thursby, G, Chintala, M, Stetson, D, Wigand, C and Champlin, D (2002) A rapid, non-destructive method for estimating aboveground biomass of salt marsh grasses. Wetlands 22(3), 626630.CrossRefGoogle Scholar
Törnqvist, TE, Cahoon, DR, Morris, JT and Day, JW (2021) Coastal wetland resilience, accelerated sea‐level rise, and the importance of timescale. Advances 2(1), e2020AV000334.Google Scholar
Tran, N and Ngo, QT (2000) Development history of deposits in the Quaternary of Vietnam. In: Nguyen, TV (ed.), The Weathering Crust and Quaternary Sediments in Vietnam. Department of Geology and Minerals of Vietnam, Hanoi, pp. 177192.Google Scholar
Vacchi, M, Engelhart, SE, Nikitina, D, Ashe, EL, Peltier, WR, Roy, K, Kopp, RE and Horton, BP (2018) Postglacial relative sea-level histories along the eastern Canadian coastline. Quaternary Science Reviews 201, 124146.CrossRefGoogle Scholar
Van Breemen, N (1982) Genesis, morphology, and classification of acid sulfate soils in coastal plains. Acid Sulfate Weathering 10, 95108.Google Scholar
Van Dam, BR, Zeller, MA, Lopes, C, Smyth, AR, Böttcher, ME, Osburn, CL, Zimmerman, T, Pröfrock, D, Fourqurean, JW and Thomas, H (2021) Calcification-driven CO2 emissions exceed “blue carbon” sequestration in a carbonate seagrass meadow. Science Advances 7(51), eabj1372.CrossRefGoogle Scholar
Van de Broek, M, Vandendriessche, C, Poppelmonde, D, Merckx, R, Temmerman, S and Govers, G (2018) Long-term organic carbon sequestration in tidal marsh sediments is dominated by old-aged allochthonous inputs in a macrotidal estuary. Global Change Biology 24(6), 24982512.CrossRefGoogle Scholar
Vilanova, I, Prieto, AR and Espinosa, M (2006) Palaeoenvironmental evolution and sea‐level fluctuations along the southeastern Pampa grasslands coast of Argentina during the Holocene. Journal of Quaternary Science: Published for the Quaternary Research Association 21(3), 227242.CrossRefGoogle Scholar
Wang, F, Lu, X, Sanders, CJ and Tang, J (2019) Tidal wetland resilience to sea level rise increases their carbon sequestration capacity in United States. Nature Communications 10(1), 111.CrossRefGoogle ScholarPubMed
Wang, F, Sanders, CJ, Santos, IR, Tang, J, Schuerch, M, Kirwan, ML, Kopp, RE, Zhu, K, Li, X and Yuan, J (2021) Global blue carbon accumulation in tidal wetlands increases with climate change. National Science Review 8(9), nwaa296.CrossRefGoogle ScholarPubMed
Wang, X, Sun, X, Wang, P and Stattegger, K (2009) Vegetation on the Sunda shelf, South China Sea, during the last glacial maximum. Palaeogeography, Palaeoclimatology, Palaeoecology 278(1–4), 8897.CrossRefGoogle Scholar
Webb, EL, Friess, DA, Krauss, KW, Cahoon, DR, Guntenspergen, GR and Phelps, J (2013) A global standard for monitoring coastal wetland vulnerability to accelerated sea-level rise. Nature Climate Change 3(5), 458465.CrossRefGoogle Scholar
West, RC (1956) Mangrove swamps of the Pacific Coast of Colombia. Annals of the Association of American Geographers 46(1), 98121.CrossRefGoogle Scholar
Williamson, P and Gattuso, J-P (2022) Carbon removal using coastal blue carbon ecosystems is uncertain and unreliable, with questionable climatic cost-effectiveness. Frontiers in Climate 4, 853666.CrossRefGoogle Scholar
Woodroffe, CD (1981) Mangrove swamp stratigraphy and Holocene transgression, grand Cayman Island, West Indies. Marine Geology 41(3), 271294.CrossRefGoogle Scholar
Woodroffe, CD and Davies, G (2009) The morphology and development of tropical coastal wetlands. In Perillo, G, Wolanski, E, Cahoon, D and Brinson, M (eds.), Coastal Wetlands: An Integrated Ecosystem Approach. Amsterdam, The Netherlands and Oxford, United Kindgdom: Elsevier, pp. 6588.Google Scholar
Woodroffe, CD, Mulrennan, ME and Chappell, J (1993) Estuarine infill and coastal progradation, Southern van Diemen Gulf, Northern Australia. Sedimentary Geology 83(3–4), 257275.CrossRefGoogle Scholar
Woodroffe, CD and Murray-Wallace, CV (2012) Sea-level rise and coastal change: The past as a guide to the future. Quaternary Science Reviews 54, 411.CrossRefGoogle Scholar
Woodroffe, CD, Thom, BG and Chappell, J (1985) Development of widespread mangrove swamps in mid-Holocene times in Northern Australia. Nature 317, 711713.CrossRefGoogle Scholar
Worthington, TA, Zu Ermgassen, PS, Friess, DA, Krauss, KW, Lovelock, CE, Thorley, J, Tingey, R, Woodroffe, CD, Bunting, P and Cormier, N (2020) A global biophysical typology of mangroves and its relevance for ecosystem structure and deforestation. Scientific Reports 10(1), 111.CrossRefGoogle ScholarPubMed
Ximenes, A, Maeda, E, Arcoverde, G and Dahdouh-Guebas, F (2016) Spatial assessment of the bioclimatic and environmental factors driving mangrove tree species’ distribution along the Brazilian coastline. Remote Sensing 8(6), 451.CrossRefGoogle Scholar
Xiong, Y, Liao, B and Wang, F (2018) Mangrove vegetation enhances soil carbon storage primarily through in situ inputs rather than increasing allochthonous sediments. Marine Pollution Bulletin 131, 378385.CrossRefGoogle ScholarPubMed
Figure 0

Figure 1. Global mapped distribution and existing estimates of carbon cycling parameters of (A) saltmarsh and (B) mangrove; (C) modelled distribution of seagrass; and (D) genus richness of benthic marine macroalgae. All values are global mean values ±1 standard deviation (where available) unless otherwise specified. Belowground carbon stocks are estimated to 1 m depth. CAR = (surface) carbon accumulation rate; NPP = net ecosystem primary productivity; SE = 1 standard error. Note that for macroalgae, CAR is replaced by estimates of carbon burial in situ (i.e. in algal beds) and exported particulate organic carbon buried in shelf sediments. Map data sources: saltmarsh (Mcowen et al., 2017); mangrove (Bunting et al., 2018); seagrass (Jayathilake and Costello, 2018); macroalgae (Kerswell, 2006). Carbon data sources: a (Rogers et al., 2019a); b (Pendleton et al. 2012a); c (Duarte and Cebrian, 1996); d (Wang et al., 2021); e (Ouyang and Lee, 2014); f (Atwood et al., 2017); g; h (Alongi, 2012); i; j (McLeod et al., 2011); k (Krause-Jensen and Duarte, 2016).

Figure 1

Figure 2. Profiles of BCE landscapes indicating (A) accommodation space, delimited by the highest astronomical tide, basement or bedrock geology, and hydrodynamic conditions favourable for mineral and organic matter accumulation (modified from Rogers (2021)); and (B) the range of techniques that can be used to observe and measure changes in substrate volume, mineral and organic matter accumulation, and position within the tidal frame, with specific focus on changing tidal position with sea-level rise, as per Allen (2000).

Figure 2

Figure 3. (A) Conceptual model of lateral zonation of BCEs of southeastern Australia with respect to tidal parameters, and varying distribution of soil organic carbon within the active root zone, inactive root zone and subtidal zone; and (B) associated generalised variation in carbon storage within BCEs (modified from Owers et al. (2022)). (C and D) Relationships between carbon storage and sea-level change (from T0 to Tn) under conditions of (C) relatively stable sea level since the mid-Holocene and (D) rising sea level since the mid-Holocene (modified from Allen (2000)).

Figure 3

Figure 4. Relative sea-level change is a significant control on processes of carbon accumulation and decomposition, and varies globally according to the generalised Holocene relative sea-level zones (A) (Clark et al., 1978) and generalised Holocene relative sea-level curves across these zones (B). Note the distribution of coastal acid sulphate soils in (Michael, 2013) (A), which corresponds broadly with regions where sea-level conditions facilitated widescale mangrove and saltmarsh development throughout the late-Holocene. When sea levels are rising (C), sedimentary carbon continues to accumulate within available accommodation and pathways of decomposition are dampened under increasingly anaerobic conditions. Where sea level has been relatively stable (D), blue carbon additionality is limited by the upper limit of tidal inundation and substrates become increasingly mineral dominated and support terrestrial vegetation as accommodation diminishes. Under conditions of falling sea levels (E), substrates become increasingly terrestrialised (i.e. with terrestrial vegetation) and conditions favour aerobic decomposition and methanogenesis of blue carbon.

Figure 4

Figure 5. Conceptualisation of the interacting effects of atmospheric carbon dioxide, warming and relative sea-level rise on BCE projected to occur under a range of emissions scenarios. Under the baseline scenario (A) carbon is fixed by in situ vegetation and contributes to soil carbon accumulation and substrate volume via accretion. The landward margin under brackish conditions is a source of methane. Under the mid-range emissions scenario (B) the feedbacks between elevated atmospheric CO2 and organic carbon sequestration and between relative sea-level rise (RSLR) and organic carbon production, preservation, and vertical accretion are strengthened. Saline intrusion reduces methane emissions in the landward fringe, although this may be counterbalanced by increased emissions resulting from increases in NPP induced by CO2 and warming. Under the high-range emission scenario (C) a tipping point is reached where RSLR exceeds vertical accretion leading to mortality and shoreline retreat. Mortality of terrestrial vegetation contributes to elevated methane production in the short term. Note that the relative strength of interactions with greenhouse gases is indicated by the thickness of lines, and greyed-out vegetation is indicative of dieback or loss associated with the effects of rising sea levels.

Supplementary material: File

Rogers et al. supplementary material

Rogers et al. supplementary material

Download Rogers et al. supplementary material(File)
File 29.8 KB

Author comment: The present, past and future of blue carbon — R0/PR1

Comments

No accompanying comment.

Review: The present, past and future of blue carbon — R0/PR2

Conflict of interest statement

No competing interests

Comments

This work is interesting and timely and provides a nice, albeit geographically biased, overview of the topic. There are however some shortcomings.

There are quite a few orthographic and spelling mistakes throughout, which should be addressed prior to resubmission, without line or page numbers it is difficult to highlight, but the first line of the abstract doesn’t make sense and should be rewritten, the other orthographic issues are on pg 2, 7, 9, 10 (a few),11, 12, 13 (a few), and 15 of the MS.

These are the minor issues.

Larger issues are that for a big review such as this there is not a lot of information on regions outside of the US, Australia and (a bit on some parts of Asia), particularly Europe and S. America, although there are a lot of articles on both SLR, and to a lesser extent, blue carbon from these areas.

On pg 4, when referring to sequestration in seagrasses, it should also be noted that seagrasses can export carbon to adjacent ecosystems including mudflats and in the final sentence of the first section on that page, it would be better to justify the focus of the paper on salt marshes and mangroves on the greater C storage capacity per area rather than solely on in situ storage.

On page 5 the second from last sentence focusses on the geographical regions of N America and Australia (plus S Africa). Is this solely because there has been a greater focus of study written in English on these regions, or potentially just studies that the authors have read on these regions. there is at least one study in S. America that discusses this, and potentially more in other regions.

On page 10, the first section (carried on from the previous page) is also focussed on particular regions. There are quite a few studies from groups in Europe (Netherlands, Spain, Germany, Fennoscandia, the Baltic States, UK, and France) and S America that are missing from this, including groups from the National University of Mar del Plata in Argentina, Marcelo Cohens group in Brazil, Luiz Drude de Lacerda’s group in another part of Brazil, and groups from Santa Catarina and Sao Paulo in Brazil that work in this area as well as Sanders from the US who works in S. America.

Lastly, the first paragraph of the conclusions finishes with a very broad generalisation of the southern hemisphere, which based on the review, really only accounts for Australia. It would be better to either broaden out the review to a wider geographical coverage or undertake a systematic review (this would be the best way forwards), that will truly account for global coverage, even if only English language publications.

Recommendation: The present, past and future of blue carbon — R0/PR3

Comments

Associate Editor (Krauss): My apologies for the time it took to received a review of this manuscript. I sent it to 22 people, and finally received 2 commitments. Only one was returned. I think this has to do with how many people you are affiliated with globally, and/or that this is a new journal not yet recognized.

This review is very nice. Most of my comments are grammatical, but please incorporate those comments and ideas provide by the reviewer. Upon revision, I only ask that you address how the reviewer’s comments were handled in your cover letter. No need to address my many minor edits.

Thank you for the submission. I look forward to reviewing the final revised version, as well as being able to cite this manuscript at a future date.

Editorial review,

Need to define “SLR” upon its first usage.

P. 4. McLeod et al., 2011. Should be a lower case “l” in name.

P. 5. Would you be willing to define “Sabkha” parenthetically next to its first usage. Not sure many readers will be familiar with that term.

P. 6. Need a period after “…Duarte 2016).”

P. 6. Please define the abbreviation “SRTM” upon first usage, or just delete the abbreviation. Next line, “or” should be “of”.

P. 7. Re: the statement “… of at least 20 years before harvesting can occur”. Also, in the tropical cyclone belt, harvesting would be replaced by storm effects over a similar period of time. Might be worth incorporating. This is our biggest point of discussion in the Neotropics where harvesting is limited. I tend to equate this cyclone destruction to annual marsh senescence; instead of an annual cycle of aboveground C turnover, mangroves can have a two-decade cycle of turnover.

P. 7. Presumption of synchronous C burial with aboveground C increment is interesting though. I feel like you have been able to document fairly well that this synchrony is not true for mangroves under rising sea-levels.

P. 7. Duplicate citation for “Saintilan 1997” on two occasions.

P. 7. Re: the statement “…where mangroves have expanded into saltmarsh zones positioned higher in the tidal frame” does not apply to the neotropics. I know that you have previously addressed this, but perhaps a caveat is needed here. Marshes are most often lower in the tidal frame than mangroves in the neotropics. You always step down when you transition from mangrove to marsh.

P. 9. Re: the statement “…oxidizing materials required for decomposition of soil organic matter.” Remember also that there are anaerobic pathways of soil organic matter decomposition. You address this immediately below, but I suggest instead of “required” consider a different word or phrase here.

P. 9. Should be “decomposition of organic matter”. Missing the word “of”.

P. 9. Line with the citation to Allen (2009); need to add space between “and” and “organic”.

P. 9. Hydroperiod also includes “depth,” or rather should be “depth, duration, and frequency”.

P. 10. Should the phrase be “sand sheets”?

P. 10. Be sure that “-1” is superscripted in 20 mm yr-1.

P. 10. A fairly useful perspective from the Caribbean basin is provided by: Sherrod CL, McMillan C (1985) The distributional history and ecology of mangrove vegetation along the northern Gulf of Mexico coastal region. Contributions in Marine Science, 28, 129–140. This might help with comments by the Reviewer as well. Also, look up what Maggie Toscano, Debra Willard, or Miriam Jones has published on the Everglades region. Might strengthen a few of these statements on Page 10.

P. 10. Please use “keep up” or ‘keep up’.

P. 11. In lieu of “1.5 o/oo,” I would standardize to “p.p.t.” for consistency.

P. 11. Should be “increases in atmospheric”. Need to add “in”.

P. 11. Paragraph break before “Presentation...”?

P. 11. Delete “,” after 3500 yr BP. Next line should be “reported to have”. Add the word “to”.

P. 12. Re: “…exhibit high soils organic carbon…”. Should this be “soil”?

P. 12. Mispelled “below-ground” as “belowgroun”.

P. 12. Line immediately after subheading. Delete “t”.

P. 13. Misspelled “livelihoods” and “security”.

P. 13. Need a space after “markets”

P. 13. Double citation of “Lovelock et al. 2022”

P. 14. Re: “…whilst the 100 year timeframe aligns with what is regarded to be “sequestered carbon”.” Is there is citation for the usage of this time frame? I’ve often wondered if there was a citation for that assumption. It seems to be more widely known in the core world as 100 years.

P. 15. Misspelled “flux” as “flx”. Next line, misspelled “atmospheric”.

P. 15. Should be “outcomes are”.

P. 16. Should this be “vary” instead of “varies”? Refers back to “magnitude and duration”. Two lines later, misspelled “between”.

P. 17. Re: “…capacity of at the site scale…”. Need to delete the word “of”. Few lines later, need space between “and” and “policy”.

P. 18. Misspelled “appreciate”. Also, should be “data were”.

Decision: The present, past and future of blue carbon — R0/PR4

Comments

No accompanying comment.

Author comment: The present, past and future of blue carbon — R1/PR5

Comments

Dear Professor Tom Spencer;

Thank you for the invitation to submit an article to Cambridge Prisms: Coastal Futures focused on Blue Carbon. We appreciate the difficulty in finding reviewers for manuscripts; however, the reviews we received have allowed us to reflect on the manuscript and improve the submission.

In the attached document we have detailed explicitly how we have addressed each comment provided by reviewers. Here we have taken the opportunity to emphasise the primary changes that have been made to the manuscript. These changes have been substantial, but we feel have improved the manuscript. We have attached both a clean and tracked changes version of this manuscript.

Both reviewers suggested expanding upon the Holocene sea-level literature pertaining to the changing distribution of mangroves and saltmarshes with relative sea-level change, and the editor-in-chief (you) suggested structuring this section to focus more on relative sea-level rise zones proposed by Clark et al. (1978). In response, we have restructured the section titled “The PAST” to include considerably more global literature; however, to maintain brevity this section may still appear far from comprehensive. We agree with reviewer 2 that a comprehensive or systematic review is needed, but this is beyond the scope of this manuscript. We disagree that there was a focus on English literature, rather the focus was on regions where sea level has been rising and where sea level has been stable – much of this literature is from countries where English is the primary language; we have added more references to include literature from other regions, including South America.

Some changes to figures have occurred. Figure 4 has been replaced with schematics indicating sea level zones by Clark et al. (1978) and the varying response of substrates and blue carbon to relative sea-level change throughout the Holocene. Specific reference to this figure is made in “The PAST” and this addition facilitated structuring this section around the Holocene relative sea-level zones demarcated by Clark et al. (1978). We have also improved Figure 5 to provide additional detail regarding carbon fluxes across wetland substrates.

Again, we appreciate the informative reviews that were received. We look forward to receiving further guidance about the manuscript.

Kind regards

Kerrylee Rogers

Recommendation: The present, past and future of blue carbon — R1/PR6

Comments

Dr. Rogers - Nice job on the revision. Sorry for my delay in accepting your manuscript. We really appreciate you writing this manuscript, which should put this message forward into many hands. Your manuscript will also help advance the stature of Coastal Futures tremendously!

Decision: The present, past and future of blue carbon — R1/PR7

Comments

No accompanying comment.