Disclaimer: This dissertation has been written by a student and is not an example of our professional work, which you can see examples of here.

Any opinions, findings, conclusions, or recommendations expressed in this dissertation are those of the authors and do not necessarily reflect the views of UKDiss.com.

Nitrous oxide and methane in a changing Arctic Ocean

Info: 12342 words (49 pages) Dissertation
Published: 31st Jan 2022

Reference this

Tagged: Environmental Science

Authors: Andrew P. Rees, Hermann W. Bange, Damian L. Are´valo-Martı´nez, Yuri Artioli, Dawn M. Ashby, Ian Brown, Hanna I. Campen, Darren R. Clark, Vassilis Kitidis, Gennadi Lessin, Glen A. Tarran, Carol Turley

Abstract

Human activities are changing the Arctic environment at an unprecedented rate resulting in rapid warming, freshening, sea ice retreat and ocean acidification of the Arctic Ocean. Trace gases such as nitrous oxide (N2O) and methane (CH4) play important roles in both the atmospheric reactivity and radiative budget of the Arctic and thus have a high potential to influence the region’s climate. However, little is known about how these rapid physical and chemical changes will impact the emissions of major climate-relevant trace gases from the Arctic Ocean. The combined consequences of these stressors present a complex combination of environmental changes which might impact on trace gas production and their subsequent release to the Arctic atmosphere. Here we present our current understanding of nitrous oxide and methane cycling in the Arctic Ocean and its relevance for regional and global atmosphere and climate and offer our thoughts on how this might change over coming decades.

Keywords: Arctic Ocean, Environmental change, Ice retreat, Methane, Nitrous oxide, Ocean acidification, Warming

BACKGROUND

The Earth’s polar regions are rapidly changing as a direct result of our altered climate. The recent IPCC special report on the oceans and the cryosphere (IPCC 2019) found that ‘‘The polar regions are losing ice, and their oceans are changing rapidly. The consequences of this polar transition extend to the whole planet, and are affecting people in multiple ways’’. The changes that are being experienced are having and will continue to have an increasing effect on the biogeochemical processes which are fundamental to the functioning of marine ecosystems (Rees 2012). The Arctic Ocean (AO) is experiencing this climate driven modification of its environment faster than anywhere else on the globe and there is high confidence that this region will likely become practically sea ice–free during the seasonal sea ice minimum for the first time before 2050, the practically ice-free state is projected to occur more often and with higher greenhouse gas concentrations (Fox-Kemper et al. 2021). To add to the complexity of this change, the loss of ice will be accompanied by the combined effects of increasing temperatures (Huang et al. 2017) and ocean acidification (AMAP 2018). Each of these artefacts of change have the potential to disrupt the biological processes which control the production and consumption of the atmospherically important trace gases nitrous oxide (N2O) and methane (CH4), whilst some of these effects may also modify the physical exchange between ocean layers or between the ocean and atmosphere.

As both of these gases are strongly radiatively active and their release from the Arctic Ocean will be impacted by ongoing and projected environmental changes, there is a need to further our understanding of how this might alter regional and global climate and ultimately impact society. The project ‘‘Pathways and emissions of climate-relevant trace gases in a changing Arctic Ocean (PETRA)’’ was designed to address these issues. In this paper we set observational and experimental evidence from a PETRA research cruise to the Fram Strait in July 2018 on board RV Polarstern (PS114) within the context of previous measurements of N2O and CH4 made by ourselves and others over the pan-Arctic region to assess our current understanding and to consider likely scenarios of change of N2O and CH4 in a future Arctic Ocean.

NITROUS OXIDE (N2O)

N2O is a long-lived atmospheric trace gas whose atmospheric mixing ratio is increasing at a mean rate of 0.85 ± 0.03 ppb yr–1 (Canadell et al. 2021). It is rated as the third most important greenhouse gas (GHG) in the troposphere (Butler and Montzka 2018; Canadell et al. 2021) with a global warming potential on a 100-year timescale of approximately 300 times that of CO2 (Ramaswamy et al. 2001; Etminan et al. 2016) and is involved in ozone (O3) depletion in the stratosphere (Ravishankara et al. 2009). Though the N2O concentration in most of the surface of the ocean is in close equilibrium with the atmosphere (Nevison et al. 1995), global emissions from the open ocean and coastal waters contribute 35–39% of the total natural sources of N2O (Tian et al. 2020) and there is a fine balance between the ocean acting as net producer or consumer of N2O. Environmental effects associated with a changing climate, which include rising temperatures, oxygen depletion and ocean acidification are quite likely to impact the level of this equilibrium (Bange et al. 2019).

Nitrous oxide is biologically produced through three processes: Denitrification is the anaerobic reduction of NO3- to N2 which has N2O as an intermediate; nitrification involves the two stage aerobic oxidation of NH4? through NO2- to NO3-, where the release of N2O as a by-product is dependent on the ambient O2 concentration (Goreau et al. 1980; Lo¨scher et al. 2012). In the third route, nitrifierdenitrification, N2O can be formed during the reduction of NO2- via nitric oxide to N2O by ammonia oxidizing bacteria. The pathway by which N2O is produced by ammonia oxidizing archaea is not yet fully understood (Wu et al. 2020).

METHANE (CH4)

CH4 is the most abundant organic trace gas in the environment, it plays an important role in the Earth’s climate and as a result of anthropogenic activities its atmospheric mixing ratio has more than doubled since the preindustrial era (Etminan et al. 2016). CH4 acts to limit the tropospheric oxidative capacity and is the second most important greenhouse gas, with a global warming potential that exceeds CO2 by up to 32 times over a 100-year timescale (Etminan et al. 2016; Canadell et al. 2021), contributing approximately 20% of the radiative climate forcing for all GHGs. The world’s oceans are a natural source of CH4 but play only a minor role in its global atmospheric budget, the open ocean and coastal waters account for 7 to 12% of the total natural sources and approximately 4% of global emissions (Saunois et al. 2020). It is thought that the open ocean contributes only a minor proportion of atmospheric CH4 with coastal environments including estuaries thought to account for approximately 75% of the total marine source (Weber et al. 2019).

The origin of CH4 in marine waters is from a diverse range of sources which may be geological, including from hydrothermal vents, cold seeps, and CH4 clathrates, or microbial in origin. The source of CH4 in ocean waters continues to prove enigmatic. Traditional understanding suggests that microbial methanogenesis is an anaerobic process, which in oceanic waters is thought to occur either in oxygen deplete waters or in anoxic micro-environments that are associated with zooplankton guts and particulate material (Brooks et al. 1981; Bianchi et al. 1992; Marty et al. 1997). However, the enhanced CH4 surface saturations found far away from shelf areas are difficult to explain because conventional CH4 production via archaeal methanogenesis should not occur in the well-oxygenated surface waters. This ‘‘ocean methane paradox’’ has been explained in tropical and sub-tropical oligotrophic conditions by the decomposition of methylphosphonate in phosphorus starved conditions (Karl et al. 2008) whereas in Arctic waters an alternative in situ CH4 production has been proposed following the microbial cleavage of dimethylsulphoniopropionate (DMSP) (Damm et al. 2008, 2010, 2015). Indeed DMSP is often found in very high concentrations in the AO during sea ice algal blooms and sea ice brines, and is projected to increase following ongoing changes to the Arctic environment (Campen et al. 2021).

CURRENT UNDERSTANDING OF N2O AND CH4 IN ARCTIC WATERS

Concentrations of dissolved N2O in the surface waters of Arctic shelf areas and central deep basins which exchange with the atmosphere are remarkably variable and can range from pronounced undersaturation to high supersaturation (Table 1). In general, ice-free surface waters appear to be undersaturated with N2O whereas ice-covered surface waters are supersaturated with N2O (Kitidis et al. 2010; Randall et al. 2012; Fenwick et al. 2017). For the North American AO it was suggested that high levels of N2O were largely associated with production in the shelf sediments of Chukchi and Bering Sea’s with subsequent advection eastwards (Fenwick et al. 2017). Supersaturations of N2O were also associated with the continental shelf for this region during the 7th Chinese National Arctic Research Expedition (Zhan et al. 2017, 2021) who also noted that offshore waters tended to be in equilibrium with the atmosphere. The Nordic Seas associated with the Greenland Basin and Fram Strait have been reported to be undersaturated in N2O (Rees et al. 2021) and act as a permanent sink to atmospheric N2O (Zhan et al. 2016).

Table 1 Exemplar N2O and CH4 publications from Arctic waters to indicate variability in regional source-sink characteristics for the two gases in the upper water column likely to exchange with the overlying atmosphere

Region

Saturaon Range (%)

(Concentraon Range (nM))

Potenal for Source - Sink

Reference

N2O

     

Labrador Sea – South Baffin Bay

Baffin Bay – Marginal ice

Baffin Bay North & Lancaster Sound

Canadian Archipelago

Amundsen Gulf – Beaufort Sea

109 ± 9 (15.4 ± 1.6)

117 ± 24 (19.8 ± 3.1)

116  ± 12 (17.6 ± 1.6)

117  ± 15 (21.0 ±2.8)

115 ± 7 (17.0 ± 1.0)

Equilibrium-Source

Source

Source

Source

Source

(Kidis et al. 2010)

<100% aributed to melt-water.

>100% found under mul-year sea-ice

Amundsen Gulf:                Sea Ice

Water under ice

~15 – 39 (<3 – 7.9) 60 – 111 (11 – 18.8)

Sink

Sink

(Randall et al. 2012)

Chukchi Sea:                Inner shelf

Outer shelf Abyssal plain

106 – 157 (12.1 – 25.1)

95 – 100 (12.7 – 16.4)

92 – 94 (16.3 – 16.9)

Source

Equilibrium-Sink

Sink

(Hirota et al. 2009, Zhan et al. 2017, Zhan et al. 2021)

(Zhang et al. 2015)

(Zhang et al. 2015, Zhan et al. 2021)

Bering Sea

102 – 137 (9.6 – 21.8)

Source

(Hirota et al. 2009, Zhan et al. 2017, Zhan et al. 2021)

North American Arcc:

Surface waters – ice free

Surface waters – ice covered

Alaskan Coastal Water

Pacific Summer Water

Pacific Winter Water

77 – 145 (10.9 – 24.6)

<100%

<110% (<17.8)

~100%

<120% (mean 17.0)

<145% (mean 24.6)

Sink

Source

Equilibrium

Source

Source

(Fenwick et al. 2017)

Eurasian basin:       Nansen Basin

Amundsen Basin

42 – <100

<100 – 111

Sink

Source

(Verdugo et al. 2016)

Nordic Seas

82 - 100 (9.7 – 15.1)

95 – 107 (11.6 – 18.2)

Sink

Sink to Source

(Zhan et al. 2016, Rees et al. 2021)

This study, Fig. 1a

CH4

     

Labrador Sea – South Baffin Bay

Baffin Bay – Marginal ice

Baffin Bay North & Lancaster Sound

Canadian Archipelago

Amundsen Gulf – Beaufort Sea

139 ± 45 (4.2 ± 1.3)

170 ± 51 (5.9 ± 1.6)

164 ± 65 (5.3 ± 2.1)

226 ± 94 (8.2 ± 3.5)

151 ± 76 (4.9 ± 2.8)

Equilibrium-Source Source

Equilibrium-Source

Source

Equilibrium-Source

(Kidis et al. 2010)

North American Arcc:

Surface Bering, Chukchi, Canada Basin.

95 – 220 (3.0 – 7.3) 97 – 517 (3.5 – 20.9)

Equilibrium-Source

(Fenwick et al. 2017)

(Li et al. 2017)

Surface Canadian Arcc Archipelago

<420% (mean 15.0)

Source

(Fenwick et al. 2017)

Central Arcc:

Atlanc Water

Pacific Water

(1.5 – 3.5)

(~4 - ~6)

~Equilibrium

Source

(Damm et al. 2010)

Fram Strait:

Atlanc Water

Pacific Water

(3  – 3.5)

(4  – 9)

~Equilibrium

Source

(Damm et al. 2015)

Nordic Seas

93 - 132

Sink to Source

This study, Fig. 1b

During research cruise PS114 onboard the German vessel RV Polarstern to the Fram Strait in July 2018 we measured concentrations and saturations of dissolved N2O and CH4 (Fig. 1) using an autonomous equilibrator headspace setup coupled to a trace gas cavity ringdown spectroscopy analyzer (Picarro Inc., USA; see Supplementary Material for methodology). The saturation of a dissolved gas indicates the relative amount of gas held relative to the value expected (100%) when the water body is in equilibrium with the overlying atmosphere. Supersaturated waters ([100%) indicate a local source of the particular gas whereas a value of undersaturation (\100%) indicates the potential for a sink to the atmosphere. In Fig. 1a, saturations of N2O can be seen to vary from a minimum of 95% in the southeast of the region close to Svalbard associated with the poleward West Spitsbergen Current and Atlantic influenced open-water. In the west of the region, higher levels of N2O, up to 107%, were found in shallower, icecovered water close to Greenland and associated with the southerly flowing East Greenland Current. The heterogeneous distribution of N2O does not relate directly to surface ice-cover and maybe associated with the origin of source waters, though the higher concentrations to the west are likely the product of shelf sediment production and limited exchange with the atmosphere.

At present, N2O emission estimates indicate that the overall net release of N2O from the AO to the atmosphere is comparably low (Fenwick et al. 2017). Microbial nitrification and/or denitrification in shelf sediments as well as water column nitrification have been proposed as the main N2O production pathways in the AO shelf areas and central deep basins (Kitidis et al. 2010; Verdugo et al. 2016; Fenwick et al. 2017), whereas loss processes include biological consumption (Verdugo et al. 2016; Rees et al. 2021) and physical advection (Zhan et al. 2016). Indications for the future release of N2O are conflicting. It was suggested that the ongoing decline of the Arctic sea ice cover may enhance future N2O emissions to the atmosphere (Kitidis et al. 2010) whilst in contrast, ongoing ocean acidification (AMAP 2018) has been shown to decrease N2O production in AO waters (Rees et al. 2016). The dominating N2O pathways and their dependence from changing environmental parameters/stressors (temperature, ocean acidification) and modifications of exchange across boundary layers need to be verified. Improved emission estimates of N2O to the atmosphere, as well as mechanistic understanding on how they might be affected by the above-mentioned perturbations is crucial since currently the uncertainties on the marine source of this gas to the atmosphere in the region are extremely high at approximately 100% (Yang et al. 2020).

Very high CH4 surface saturations have been observed in the shelf areas and the central deep basins of the AO (Table 1). Vast areas of the AO seafloor, particularly those associated with Siberian, Chukchi, and Beaufort Seas, is rich in permafrost (a potential substrate for methanogenesis) and CH4 hydrates (Chen et al. 2021). At present, the AO is a potentially important source of atmospheric CH4 (Kort et al. 2012) with sedimentary production via methanogenesis, dissociating gas hydrates and diffusion from geological dissolution providing the dominant source in shelf sea areas. Release of bubbles from sedimentary origin during the melting season represents a considerable, yet spatially constrained, source to the atmosphere (Thornton et al. 2020), though there is evidence to indicate that CH4 released at the sea bed may not reach the atmosphere, e.g. (Myhre et al. 2016). In Fig. 1b, CH4 saturations in the Fram Strait region during July 2018 can be seen to be highly variable. As with the distribution of N2O (Fig. 1a) and patterns of heterogeneity previously reported, minimum saturations (* 93%) of CH4 were associated with Atlantic influenced ice-free waters and the highest observed (132%) were found in ice-covered areas close to the Greenland coast suggesting a strong source of CH4 to the air for this region.

Future CH4 emissions from open ocean regions of the AO will largely be determined by aerobic CH4 oxidation in the water column and ‘non-conventional’ microbial CH4 production via DMSP (Damm et al. 2010, 2015). The increased supply of organic matter from rivers and permafrost thaw may further enhance microbial methanogenesis. The effect of future environmental stressors such as warming and pH on aerobic CH4 oxidation is largely unknown (James et al. 2016).

EVIDENCE FOR POTENTIAL CHANGE

Impact of warming

Warming conditions are likely to impart direct and in-direct effects on the processes controlling production and consumption of both gases. Whereas the solubility of gases is governed by well constrained laws of physics under the control of temperature and salinity, so that warmer waters hold less gas than cold ones, the composition of microbial communities and their ecological function of changing systems is much less predictable. Warming would further enhance microbial methanogenesis with a 2 increase in temperature resulting in a 25–200% increase in methanogenesis (Bange et al. 1998). Finally, sea ice loss as a result of warming may result in shorter residence time for CH4 and N2O in the water and thereby evasion to the atmosphere rather than in-water microbial processing (sea ice is currently considered as a semi-permeable barrier to air-sea exchange) (Kitidis et al. 2010). Increasing seawater temperatures lower the solubility of CH4 in seawater and allow a shallowing of the CH4 hydrate stability zone. Therefore, a small increase in seawater temperature could potentially lead to hydrate dissolution and the subsequent release of CH4 from the AO to the atmosphere (Kitidis 2009).

Fig. 1 Distribution of (a) N2O and (b) CH4 saturation during research cruise PS114 onboard the RV Polarstern during July 2018. The green dotted line in each image represents the southern extent of sea ice at 90% cover (see Supplementary Material for method description)

There is some confidence that phytoplankton primary productivity is likely to increase as ice-cover retreats (Lannuzel et al. 2020) which will increase the sediment load of organic material that is potentially used in remineralisation processes in the generation of both N2O and CH4. Increases in primary production can be attributed to the interplay between two factors: first, the increased spatial and temporal extent of open waters, and second, the enhanced nutrient input brought about by mixing, upwelling and lateral advection, all of which is fostered by increased inflow from subpolar seas (the so-called Atlantification of Arctic waters) and more frequent storm events (Polyakov et al. 2017).

As ocean temperatures rise, CH4 hydrates may become unstable releasing vast quantities of CH4 to the atmosphere which in turn may lead to further temperature increase and hydrate de-stabilisation according to the ‘‘clathrate gun’’ hypothesis (Kennett et al. 2003).

Although CH4 solubility will decrease with increasing temperature, methanotrophy will also increase (YvonDurocher et al. 2014). Methanotrophy is the dominant oceanic CH4 sink and is a first order process with respect to CH4 concentration and inversely related to its turnover time. This suggests that an enhanced sedimentary release, potentially caused by warming triggered gas hydrate dissociation may result in shorter turnover times of dissolved CH4 in the water column (James et al. 2016). However, direct ebullition from dissociating gas hydrates or thawing permafrost may result in substantial emission of CH4 to the atmosphere as bubbles rise faster than they are consumed by microbes (Shakhova et al. 2010). Decreased sea ice would also reduce the extent of areas where bubbles released from the sediments are trapped. The transfer of these bubbles across the sea surface during the melting season might represent a spatially variable, yet considerable source of CH4 to the atmosphere (Zhou et al. 2014).

IMPACT OF OCEAN ACIDIFICATION

In the open oceans nitrification is the dominant mechanism for the production of N2O. A number of studies have shown that nitrification rate is inhibited by decreasing pH (Beman etal.2010;Kitidisetal.2011)buttheimpactonN2Oappears equivocal. During experiments performed in Arctic and Antarctic waters (Rees et al. 2016) showed that whilst the microbial community of ammonia oxidising archaea (the dominant nitrifying organisms) seemed unaffected by changing pH, the production of N2O decreased at all stations by between 2.4 and 44% when pH was reduced at between values of 0.06 and 0.4 pH units. The reduction in N2O yield from nitrification was directly related to a decrease of between 28 and 67% in available NH3 as a result of the pH driven shift in the NH3:NH4? equilibrium. In the subarctic western North Pacific (Breider et al. 2019) found that decreasing the pH during experimental manipulations acted to significantly increase N2O production, whilst rates of nitrification either remained stable or decreased, indicating a de-coupling of the two processes. The differences between the findings of Rees et al (2016) and Breider et al (2019) maybe attributable to regionally associated differences in N2O production pathways, or to the relative sensitivities of microbial communities found in the two regions.

During the current project we performed four experiments during PS114 in the Fram Strait region similar to those described in Rees et al (2016). These were conducted to examine the impact of ocean acidification in isolation and in combination with warming of 2 C (see Supplementary Material for methodology). Initial findings can be seen in Fig. 2. It would appear that there is a decrease in both N2O concentration and in nitrification rate with changes of pH which were made to match future conditions indicated by representative concentration pathway (RCP) of 6.0 and 8.5. Whilst there is some inherent variability in the responses, it would appear that there is no obvious response of N2O production or nitrification rate to warming of 2 C.

To date we are unaware of experimental evidence to suggest that either methanogenesis or methanotrophy show any sensitivity to changing conditions of ocean acidification. A limited number of experiments that we have performed have all indicated that both CH4 production and consumption processes are likely to prove resilient to ocean acidification in the AO.

ECOSYSTEM MODELLING

The incorporation of up-to-date knowledge of N2O and CH4 dynamics in the marine environment into coupled physical-biogeochemical models is important to (a) expand our understanding and to test hypotheses related to the dynamics of these gases, (b) to explain spatial and temporal distribution patterns, and (c) to predict future change under the impact of multiple stressors. However there is currently insufficient understanding of the production and consumption pathways of these gases and their environmental controls which remains a limiting factor for their wider inclusion into process-based model studies, both in the AO and elsewhere.

A range of model formulations of N2O dynamics based on observational data have been developed in recent years and applied at a global scale, e.g. (Martinez-Rey et al. 2015; Ji et al. 2018) thus covering, but not focusing on, the AO region. However, implementation of regional-scale models of N2O is hindered by the scarcity of observational data and incompleteness in understanding of various pathways and their response to stressors, especially within high-gradient environments.

The issue is even more apparent for CH4, as many unknowns related to its production and consumption pathways still limit the rare modelling efforts to sensitivity studies, e.g. (Wa˚hlstro¨m and Meier 2014) who focused on the Laptev Sea region. Better understanding of the controls over the CH4 pathways is critical to constrain processbased models: for instance, the reported range of oceanic aerobic CH4 oxidation rates spans several orders of magnitude. In the AO the uncertainties in biological pathways are further augmented by a range of shelf sea processes related to permafrost thawing and CH4 hydrate dissolution, that all require advancements in understanding and consideration in the models.

The implementation of sophisticated process-based models which are able to project future emissions of N2O and CH4 under the influence of multiple stressors is severely hampered by sparse sampling (Weber et al. 2019) and lack of experimental evidence to advance statistically sound mechanistic understanding of the controlling processes. There is real need for an increased capacity of measurements such as those reported in this study to further develop this on both regional and global scales.

Fig. 2 The impact of decreasing ocean pH and increasing temperature by 2 C on (a) the rate of nitrification and (b) N2O concentration at four positions in the Fram Straits region of the AO during research cruise PS114 in July 2018 (see Supplementary Material for method description)

SOCIAL AND POLICY IMPLICATIONS

N2O and CH4 contribute significantly to climate change. They are relevant to the United Nations Framework Convention on Climate Change (UNFCCC), the primary international, intergovernmental forum for negotiating the global response to climate change. The ultimate objective of the UNFCCC is to stabilize GHG concentrations ‘‘at a level that would prevent dangerous anthropogenic (human induced) interference with the climate system’’ (UNFCCC 1992). The goal of the Paris Agreement (UN 2015), a legally binding international treaty on climate change, adopted by 196 Parties at UNFCCC COP 21 in Paris in 2015, is ‘‘to limit global warming to well below 2, preferably to 1.5 C, compared to pre-industrial levels’’. To achieve this long-term temperature goal, countries aim to reach global peaking of GHG emissions as soon as possible to realize a climate neutral world by mid-century.

UN Member States adopted 17 Sustainable Development Goals (SDGs), as part of the 2030 Agenda for Sustainable Development (United Nations 2015), a global partnership for sustainable improvement of human lives whilst protecting the environment, including oceans, and tackling climate change. The oceanic production of N2O and CH4 addressed in this study is relevant to climate change goal (SDG13), to ‘‘Take urgent action to combat climate change and its impacts’’.

The generation of impact is integral to PETRA (Fig. 3). Data generated will be managed and archived at our local oceanographic data centres (British Oceanographic Data Centre for UK and PANGEA for Germany) according to our data management plan.

Fig. 3 Theory of Change visualisation of social and policy implications associated with the PETRA project investigations of N2O and CH4 in changing AO

To ensure greater visibility and access, CO2 data collected during fieldwork expeditions in 2018 and 2019 have been submitted to the Surface Ocean CO2 Atlas (SOCAT) and included in the annual Global Carbon Project budget (Friedlingstein et al. 2020) whilst N2O and CH4 data will be submitted to the marine methane and nitrous oxide database (MEMENTO) (https://memento.geomar.de/). Outputs will also be made available through engagement with the GOA-ON observational network. ‘‘The polar regions are losing ice, and their oceans are changing rapidly. The consequences of this polar transition extend to the whole planet, and are affecting people in multiple ways’’ (IPCC 2019). Understanding the role of the ocean in the cycling and production of N2O and CH4 and how these may change in an ocean undergoing rapid and long-term change is therefore essential to the Convention and climate negotiations. The findings of our previous work (Rees et al. 2016) and the newer observations from PETRA indicate that ocean acidification has the potential to decrease N2O emissions by up to 0.8 Tg N yr-1 which is comparable to all current N2O production from fossil fuel combustion and industrial processes of 0.7 Tg N yr-1.

Table 2 Sources and sinks of N2O and CH4 in the AO, the expected effect of ocean warming (and associated melting) and acidification, as well as the level of uncertainties in the current state of knowledge

Processes Source/sink Estimated overall effecta Level of uncertainty
    Warming Ocean acidification  

N2O

Nitrification

Source

Medium

Denitrification

Source/sink

?

Medium

Air-sea exchange

Source/sink

High

Within-ice cycling & fluxes

Source/sink

?

?

Very high

CH4

Methanogenesis

Source

?

Medium

Methanotrophy

Sink

?

Medium

Aerobic oxidation

Sink

?

High

Air-sea exchange

Source/sink

?

High

Within-ice cycling & fluxes

Source/sink

?

?

Very high

a ↑ increase, ↓ decrease, ? unknown

There is some potential therefore for decreases in N2O release to the atmosphere to offer a negative feedback to global warming, though it is still too early to say as other contrasting (Breider et al. 2019) and compounding effects are still to be accounted for.

OUTLOOK

Whilst estimates of the global marine source of N2O and CH4 to the atmosphere have significantly improved (Wilson et al. 2020), a mechanistic understanding of the causes for the observed variability in sink-source dynamics and sea/ice-air gradients in the AO is missing. This degree of uncertainty is highlighted in Table 2, where current understanding of the processes controlling these gases in concert with ocean acidification and warming are presented. Reduced sea ice coverage in the AO and the adjacent subpolar regions with future warming will likely expose larger ocean surfaces to direct exchange with the atmosphere, increasing the overall source of N2O and CH4 (a positive feedback on GHG-driven warming). However, the sea air transfer depends strongly on the pre-existing gradients between both reservoirs. The balance is a complex product of several processes. During the freezing period, brine rejection leads to gas enrichment and densitydriven fluxes towards the underlying water column, but also potentially supports temporary fluxes towards the atmosphere. In contrast, during melting, a dilution effect causes gas undersaturation with respect to atmospheric equilibrium together with a strong salinity driven stratification. This would result in a net ocean uptake unless a strong source (e.g. sedimentary CH4 release) or buoyancydriven flow breaks the stratification causing deep mixing and upwelling. At present it is challenging to establish which process is dominant over the annual cycle since the cycling of N2O and CH4 within the sea ice is not well understood and local production can be masked by lateral advection. Moreover, it is not clear to what extent an ocean acidification driven change of N2O production would contribute to offset the expected increase in the emissions to the atmosphere, and whether any synergistic effects may arise.

Resolving gradients across the sea–ice–air interfaces and their spatial and temporal variability requires a combination of dedicated, multidisciplinary surveys, time series observations and the use of novel methods and autonomous platforms suitable for different ice conditions, e.g. (Bange et al. 2019; Lee et al. 2019). Studies investigating dynamics of N2O and CH4 within sea ice are scarce and therefore should be addressed in future joint projects. Recent observations from the MOSAiC expedition (https://mosaic-expedition.org/) and the Synoptic Arctic Survey (https://synopticarcticsurvey.w.uib.no/) are expected to provide important contributions towards a better understanding of the cycling and emissions of N2O and CH4 in the AO. Overall, strengthening observational capabilities in the AO will reduce the current emission uncertainties and thereby improve our projections of future GHG emission trends within the context of global coupled models.

Acknowledgements: This PETRA project was supported by the Changing Arctic Ocean programme which was co-funded by UK Natural Environment Research Council and the German Federal Ministry of Education and Research with awards NE/R012830/1 and FKZ 03F0808A, respectively.

Open Access: This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Originally published at: Ambio https://doi.org/10.1007/s13280-021-01633-8

REFERENCES

AMAP. 2018. AMAP Assessment 2018: Arctic ocean acidification. Tromsø, Norway: Arctic Monitoring and Assessment Programme. Bange, H.W., D.L. Arevalo-Martinez, M. de la Paz, L. Farias, J. Kaiser, A. Kock, C.S. Law, A.P. Rees, et al. 2019. A Harmonized nitrous oxide (N2O) ocean observation network for the 21st century. Frontiers in Marine Science. https://doi.org/10.3389/fmars.2019.00157.

Bange, H.W., S. Dahlke, R. Ramesh, L.A. Meyer-Reil, S. Rapsomanikis, and M.O. Andreae. 1998. Seasonal study of methane and nitrous oxide in the coastal waters of the Southern Baltic Sea. Estuarine, Coastal and Shelf Science 47: 807–817. https://doi.org/10.1006/ecss.1998.0397.

Beman, J.M., C.-E. Chow, A.L. King, Y. Feng, J.A. Fuhrman, A. Andersson, N.R. Bates, B.N. Popp, et al. 2010. Global declines in oceanic nitrification rates as a consequence of ocean acidification. Proceedings of the National Academy of Sciences of the United States of America 108: 208–213. https://doi.org/10.1073/pnas.1011053108.

Bianchi, M., D. Marty, J.L. Teyssie, and S.W. Fowler. 1992. Strictly aerobic and anaerobic bacteria associated with sinking particulate matter and zooplankton fecal pellets. Marine Ecology Progress Series 88: 55–60. https://doi.org/10.3354/meps088055.

Breider, F., C. Yoshikawa, A. Makabe, S. Toyoda, M. Wakita, Y. Matsui, S. Kawagucci, T. Fujiki, et al. 2019. Response of N2O production rate to ocean acidification in the western North Pacific. Nature Climate Change. https://doi.org/10.1038/s41558019-0605-7.

Brooks, J.M., D.F. Reid, and B.B. Bernard. 1981. Methane in the upper water column of the Northwestern Gulf of Mexico. Journal of Geophysical Research-Oceans 86: 1029–1040. https://doi.org/10.1029/JC086iC11p11029.

Butler, J.H., and S.A. Montzka. 2018. The NOAA Annual Greenhouse Gas Index (AGGI). National Oceanic and Atmospheric Administration. Retrieved 23 April 2021, from http://www.esrl.noaa.gov/gmd/aggi/aggi.html (Web material)

Campen, H.I., D.L. Are´valo-Martı´nez, Y. Artioli, I.J. Brown, V. Kitidis, G. Lessin, A.P. Rees, and H.W. Bange. 2021. The role of a changing Arctic Ocean and climate for the biogeochemical cycling of dimethyl sulphide (DMS) and carbon monoxide (CO). Ambio. https://doi.org/10.1007/s13280-021-01612-z.

Canadell, J.G., P.M.S. Monteiro, M.H. Costa, L. Cotrim da Cunha, P.M. Cox, A.V. Eliseev, S. Henson, M. Ishii, et al. 2021. Global carbon and other biogeochemical cycles and feedbacks. In Climate change 2021: The physical science basis. Contribution of working group i to the sixth assessment report of the intergovernmental panel on climate change, ed. V. MassonDelmotte, P. Zhai, A. Pirani, S.L. Connors, C. Pe´an, S. Berger, N. Caud, Y. Chen, et al. Cambridge/New York, NY: Cambridge University Press.

Chen, M.L., J.H. Kim, Y.K. Lee, D.H. Lee, Y.K. Jin, and J. Hur. 2021. Subsea permafrost as a potential major source of dissolved organic matter to the East Siberian Arctic Shelf. Science of the Total Environment. https://doi.org/10.1016/j.scitotenv.2021.146100.

Damm, E., E. Helmke, S. Thoms, U. Schauer, E. Nothig, K. Bakker, and R.P. Kiene. 2010. Methane production in aerobic oligotrophic surface water in the central Arctic Ocean. Biogeosciences 7: 1099–1108. https://doi.org/10.5194/bg-7-1099-2010.

Damm, E., R.P. Kiene, J. Schwarz, E. Falck, and G. Dieckmann. 2008. Methane cycling in Arctic shelf water and its relationship with phytoplankton biomass and DMSP. Marine Chemistry 109: 45–59. https://doi.org/10.1016/j.marchem.2007.12.003.

Damm, E., S. Thoms, A. Beszczynska-Moller, E.M. Nothig, and G. Kattner. 2015. Methane excess production in oxygen-rich polar water and a model of cellular conditions for this paradox. Polar Science 9: 327–334. https://doi.org/10.1016/j.polar.2015.05.001.

Etminan, M., G. Myhre, E.J. Highwood, and K.P. Shine. 2016. Radiative forcing of carbon dioxide, methane, and nitrous oxide: A significant revision of the methane radiative forcing. Geophysical Research Letters 43: 12614–12623. https://doi.org/10.1002/2016gl071930.

Fenwick, L., D. Capelle, E. Damm, S. Zimmermann, W.J. Williams, S. Vagle, and P.D. Tortell. 2017. Methane and nitrous oxide distributions across the North American Arctic Ocean during summer, 2015. Journal of Geophysical Research-Oceans 122: 390–412. https://doi.org/10.1002/2016jc012493.

Fox-Kemper, B., H.T. Hewitt, C. Xiao, G. Aðalgeirsdo´ttir, S.S. Drijfhout, T.L. Edwards, N.R. Golledge, M. Hemer, et al. 2021. Ocean, cryosphere and sea level change. In Climate change 2021: The physical science basis. Contribution of working group I to the sixth assessment report of the intergovernmental panel on climate change, ed. V. Masson-Delmotte, P. Zhai, A. Pirani, S.L. Connors, C. Pe´an, S. Berger, N. Caud, Y. Chen, et al. Cambridge/New York, NY: Cambridge University Press.

Friedlingstein, P., M. O’Sullivan, M.W. Jones, R.M. Andrew, J. Hauck, A. Olsen, G.P. Peters, W. Peters, et al. 2020. Global carbon budget 2020. Earth System Science Data 12: 3269–3340. https://doi.org/10.5194/essd-12-3269-2020.

Goreau, T.J., W.A. Kaplan, S.C. Wofsy, M.B. McElroy, F.W. Valois, and S.W. Watson. 1980. Production of NO2- and N2O by nitrifying bacteria at reduced concentrations of oxygen. Applied and Environmental Microbiology 40: 526–532.

Hirota, A., A. Ijiri, D.D. Komatsu, S.B. Ohkubo, F. Nakagawa, and U. Tsunogai. 2009. Enrichment of nitrous oxide in the water columns intheareaoftheBeringandChukchiSeas.MarineChemistry116: 47–53. https://doi.org/10.1016/j.marchem.2009.09.001.

Huang, J.B., X.D. Zhang, Q.Y. Zhang, Y.L. Lin, M.J. Hao, Y. Luo, Z.C. Zhao, Y. Yao, et al. 2017. Recently amplified arctic warming has contributed to a continual global warming trend. Nature Climate Change 7: 875–879. https://doi.org/10.1038/s41558-017-0009-5.

IPCC. 2019. Summary for policymakers. In IPCC special report on the ocean and cryosphere in a changing climate, ed. H.-O. Po¨rtner, D.C. Roberts, V. Masson-Delmotte, P. Zhai, M. Tignor, E. Poloczanska, K. Mintenbeck, M. Nicolai, et al., 3–35. Geneva: Intergovernmental Panel on Climate Change.

James, R.H., P. Bousquet, I. Bussmann, M. Haeckel, R. Kipfer, I. Leifer, H. Niemann, I. Ostrovsky, et al. 2016. Effects of climate change on methane emissions from seafloor sediments in the Arctic Ocean: A review. Limnology and Oceanography 61: S283–S299. https://doi.org/10.1002/lno.10307.

Ji, Q., E. Buitenhuis, P. Suntharalingam, J.L. Sarmiento, and B.B. Ward. 2018. Global nitrous oxide production determined by oxygen sensitivity of nitrification and denitrification. Global Biogeochemical Cycles 32: 1790–1802. https://doi.org/10.1029/2018GB005887.

Karl, D.M., L. Beversdorf, K.M. Bjorkman, M.J. Church, A. Martinez, and E.F. Delong. 2008. Aerobic production of methane in the sea. Nature Geoscience 1: 473–478. https://doi.org/10.1038/ngeo234.

Kennett, J.P., K.G. Cannariato, I.L. Hendy, and R.J. Behl. 2003. Methane hydrates in quaternary climate change: The clathrate gun hypothesis. In Methane hydrates in quaternary climate change: The clathrate gun hypothesis, ed. J.P. Kennett, K.G. Cannariato, I.L. Hendy, and R.J. Behl, 1–9. Washington, DC: American Geophysical Union.

Kitidis, V. 2009. Chapter 10. Methane biogeochemistry and carbon stores in the Arctic Ocean: Hydrates and permafrost. In Carbon capture: Sequestration and storage, ed. R.E. Hester and R.M. Harrison, 285–300. London: Royal Society of Chemistry.

Kitidis, V., B. Laverock, L.C. McNeill, A. Beesley, D. Cummings, K. Tait, M.A. Osborn, and S. Widdicombe. 2011. Impact of ocean acidification on benthic and water column ammonia oxidation. Geophysical Research Letters 38: L21603. https://doi.org/10.1029/2011GL049095.

Kitidis, V., R.C. Upstill-Goddard, and L.G. Anderson. 2010. Methane and nitrous oxide in surface water along the North-West Passage, Arctic Ocean. Marine Chemistry 121: 80–86. https://doi.org/10.1016/j.marchem.2010.03.006.

Kort, E.A., S.C. Wofsy, B.C. Daube, M. Diao, J.W. Elkins, R.S. Gao, E.J. Hintsa, D.F. Hurst, et al. 2012. Atmospheric observations of Arctic Ocean methane emissions up to 82 north. Nature Geoscience 5: 318–321. https://doi.org/10.1038/ngeo1452.

Lannuzel, D., L. Tedesco, M. van Leeuwe, K. Campbell, H. Flores, B. Delille, L. Miller, J. Stefels, et al. 2020. The future of Arctic seaice biogeochemistry and ice-associated ecosystems. Nature Climate Change 10: 983–992. https://doi.org/10.1038/s41558020-00940-4.

Lee, C.M., S. Starkweather, H. Eicken, M.-L. Timmermans, J. Wilkinson, S. Sandven, D. Dukhovskoy, S. Gerland, et al. 2019. A framework for the development, design and implementation of a sustained Arctic Ocean observing system. Frontiers in Marine Science 6: 451. https://doi.org/10.3389/fmars.2019.00451.

Li, Y.H., L.Y. Zhan, J.X. Zhang, L.Q. Chen, J.F. Chen, and Y.P. Zhuang. 2017. A significant methane source over the Chukchi Sea shelf and its sources. Continental Shelf Research 148: 150–158. https://doi.org/10.1016/j.csr.2017.08.019.

Lo¨scher, C.R., A. Kock, M. Koenneke, J. LaRoche, H.W. Bange, and R.A. Schmitz. 2012. Production of oceanic nitrous oxide by ammonia-oxidizing archaea. Biogeosciences 9: 2419–2429. https://doi.org/10.5194/bg-9-2419-2012.

Martinez-Rey, J., L. Bopp, M. Gehlen, A. Tagliabue, and N. Gruber. 2015. Projections of oceanic N2O emissions in the 21st century using the IPSL Earth system model. Biogeosciences 12: 4133–4148. https://doi.org/10.5194/bg-12-4133-2015.

Marty, D., P. Nival, and W.D. Yoon. 1997. Methanoarchaea associated with sinking particles and zooplankton collected in the northeastern tropical Atlantic. Oceanologica Acta 20: 863–869.

Myhre, C.L., B. Ferre, S.M. Platt, A. Silyakova, O. Hermansen, G. Allen, I. Pisso, N. Schmidbauer, et al. 2016. Extensive release of methane from Arctic seabed west of Svalbard during summer 2014 does not influence the atmosphere. Geophysical Research Letters 43: 4624–4631. https://doi.org/10.1002/2016gl068999.

Nevison, C.D., R.F. Weiss, and D.J. Erickson. 1995. Global oceanic emissions of nitrous-oxide. Journal of Geophysical ResearchOceans 100: 15809–15820. https://doi.org/10.1029/95JC00684.

Polyakov, I.V., A.V. Pnyushkov, M.B. Alkire, I.M. Ashik, T.M. Baumann, E.C. Carmack, I. Goszczko, J. Guthrie, et al. 2017. Greater role for Atlantic inflows on sea-ice loss in the Eurasian Basin of the Arctic Ocean. Science. https://doi.org/10.1126/science.aai8204.

Ramaswamy, V., O. Boucher, J. Haigh, D. Hauglustaine, J. Haywood, G. Myhre, T. Nakajima, G.Y. Shi, et al. 2001. Radiative forcing of climate change. In Climate Change 2001: The scientific basis. Contribution of working group I to the third assessment report of the intergovernmental panel on climate change, ed. J.T. Houghton, Y. Ding, D.J. Griggs, M. Noguer, P.J. Van der Linden, X. Dai, K. Maskell, and C.A. Johnson, 349–416. Cambridge: Cambridge University Press.

Randall, K., M. Scarratt, M. Levasseur, S. Michaud, H.X. Xie, and M. Gosselin. 2012. First measurements of nitrous oxide in Arctic sea ice. Journal of Geophysical Research-Oceans. https://doi.org/10. 1029/2011jc007340.

Ravishankara, A.R., J.S. Daniel, and R.W. Portmann. 2009. Nitrous oxide (N2O): The dominant ozone-depleting substance emitted in the 21st century. Science 326: 123–125. https://doi.org/10.1126/science.1176985.

Rees, A.P. 2012. Pressures on the marine environment and the changing climate of ocean biogeochemistry. Philosophical Transactions of the Royal Society a-Mathematical Physical and Engineering Sciences 370: 5613–5635. https://doi.org/10.1098/rsta.2012.0399.

Rees, A.P., I.J. Brown, A. Jayakumar, G. Lessin, P.J. Somerfield, and B.B. Ward. 2021. Biological nitrous oxide consumption in oxygenated waters of the high latitude Atlantic Ocean. Communications Earth & Environment 2: 36. https://doi.org/10.1038/s43247-021-00104-y.

Rees, A.P., I.J. Brown, A. Jayakumar, and B.B. Ward. 2016. The inhibition of N2O production by ocean acidification in cold temperate and polar waters. Deep-Sea Research Part Ii-Topical Studies in Oceanography 127: 93–101. https://doi.org/10.1016/j.dsr2.2015.12.006.

Saunois, M., A.R. Stavert, B. Poulter, P. Bousquet, J.G. Canadell, R.B. Jackson, P.A. Raymond, E.J. Dlugokencky, et al. 2020. The Global Methane Budget 2000–2017. Earth System Science Data 12: 1561–1623. https://doi.org/10.5194/essd-12-1561-2020.

Shakhova, N., I. Semiletov, A. Salyuk, V. Yusupov, D. Kosmach, and O. Gustafsson. 2010. Extensive methane venting to the atmosphere from sediments of the East Siberian Arctic Shelf. Science 327: 1246–1250. https://doi.org/10.1126/science.1182221.

Thornton, B.F., J. Prytherch, K. Andersson, I.M. Brooks, D. Salisbury, M. Tjernstro¨m, and P.M. Crill. 2020. Shipborne eddy covariance observations of methane fluxes constrain Arctic sea emissions. Science Advances. https://doi.org/10.1126/sciadv.aay7934.

Tian, H.Q., R.T. Xu, J.G. Canadell, R.L. Thompson, W. Winiwarter, P. Suntharalingam, E.A. Davidson, P. Ciais, et al. 2020. A comprehensive quantification of global nitrous oxide sources and sinks. Nature. https://doi.org/10.1038/s41586-020-2780-0.

Verdugo, J., E. Damm, P. Snoeijs, B. Diez, and L. Farias. 2016. Climate relevant trace gases (N2O and CH4) in the Eurasian Basin (Arctic Ocean). Deep-Sea Research Part I-Oceanographic Research Papers 117: 84–94. https://doi.org/10.1016/j.dsr.2016.08.016.

Wa˚hlstro¨m, I., and H.E.M. Meier. 2014. A model sensitivity study for the sea–air exchange of methane in the Laptev Sea, Arctic Ocean. Tellus b: Chemical and Physical Meteorology 66: 24174. https://doi.org/10.3402/tellusb.v66.24174.

Weber, T., N.A. Wiseman, and A. Kock. 2019. Global ocean methane emissions dominated by shallow coastal waters. Nature Communications 10: 4584. https://doi.org/10.1038/s41467-01912541-7.

Wilson, S.T., A.N. Al-Haj, A. Bourbonnais, C. Frey, R.W. Fulweiler, J.D. Kessler, H.K. Marchant, J. Milucka, et al. 2020. Ideas and perspectives: A strategic assessment of methane and nitrous oxide measurements in the marine environment. Biogeosciences 17: 5809–5828. https://doi.org/10.5194/bg-17-5809-2020.

Wu, L., X. Chen, W. Wei, Y. Liu, D. Wang, and B.-J. Ni. 2020. A critical review on nitrous oxide production by ammonia-oxidizing archaea. Environmental Science & Technology 54: 9175–9190. https://doi.org/10.1021/acs.est.0c03948.

Yang, S., B.X. Chang, M.J. Warner, T.S. Weber, A.M. Bourbonnais, A.E. Santoro, A. Kock, R.E. Sonnerup, et al. 2020. Global reconstruction reduces the uncertainty of oceanic nitrous oxide emissions and reveals a vigorous seasonal cycle. Proceedings of the National Academy of Sciences 117: 11954. https://doi.org/10.1073/pnas.1921914117.

Yvon-Durocher, G., A.P. Allen, D. Bastviken, R. Conrad, C. Gudasz, A. St-Pierre, N. Thanh-Duc, and P.A. del Giorgio. 2014. Methane fluxes show consistent temperature dependence across microbial to ecosystem scales. Nature 507: 488–491. https://doi.org/10.1038/nature13164.

Zhan, L.Y., L.Q. Chen, J.X. Zhang, J.P. Yan, Y.H. Li, and M. Wu. 2016. A permanent N2O sink in the Nordic Seas and its strength and possible variability over the past four decades. Journal of Geophysical Research-Oceans 121: 5608–5621. https://doi.org/10.1002/2016jc011925.

Zhan, L.Y., M. Wu, L.Q. Chen, J.X. Zhang, Y.H. Li, and J. Liu. 2017. The air-sea nitrous oxide flux along cruise tracks to the Arctic Ocean and Southern Ocean. Atmosphere 8: 216. https://doi.org/10.3390/atmos8110216.

Zhan, L.Y., J.X. Zhang, Z.X. Ouyang, R.B. Lei, S.Q. Xu, D. Qi, Z.Y. Gao, H. Sun, et al. 2021. High-resolution distribution pattern of surface water nitrous oxide along a cruise track from the Okhotsk Sea to the western Arctic Ocean. Limnology and Oceanography 66: S401–S410. https://doi.org/10.1002/lno.11604.

Zhang, J., L. Zhan, L. Chen, Y. Li, and J. Chen. 2015. Coexistence of nitrous oxide undersaturation and oversaturation in the surface and subsurface of the western Arctic Ocean. Journal of Geophysical Research: Oceans 120: 8392–8401. https://doi.org/10.1002/2015JC011245.

Zhou, J., J.L. Tison, G. Carnat, N.X. Geilfus, and B. Delille. 2014. Physical controls on the storage of methane in landfast sea ice. The Cryosphere 8: 1019–1029. https://doi.org/10.5194/tc-81019-2014.

AUTHOR BIOGRAPHIES

Andrew P. Rees (&) is a senior scientist at the Plymouth Marine Laboratory. His research interests include the biogeochemical cycling of carbon and nutrients and the ocean-atmosphere interaction of greenhouse gases in estuaries, coastal waters and the open ocean. Address: Plymouth Marine Laboratory, Prospect Place, The Hoe, Plymouth PL1 3DH, UK. e-mail: apre@pml.ac.uk

Hermann W. Bange is a Professor at the GEOMAR Helmholtz Centre for Ocean Research Kiel, Germany. His research interests include trace gas biogeochemistry and the nitrogen and sulphur cycles in the open and coastal oceans. Address: GEOMAR Helmholtz-Zentrum Fu¨r Ozeanforschung Kiel, Chemische Ozeanographie, Du¨sternbrooker Weg 20, 24105 Kiel, Germany. e-mail: hbange@geomar.de

Damian L. Are´valo-Martı´nez is a postdoctoral researcher at the Institute of Geosciences of Kiel University and the Chemical Oceanography Department of GEOMAR Helmholtz Centre for Ocean Research Kiel. His research interests include biogeochemistry of climate-relevant trace gases, gas exchange across sea-air and sea-ice interfaces, laser spectroscopy-based methods for trace gas analysis, time-series observations, and land-ocean connectivity through groundwater. Address: GEOMAR Helmholtz-Zentrum Fu¨r Ozeanforschung Kiel, Chemische Ozeanographie, Du¨sternbrooker Weg 20, 24105 Kiel, Germany. e-mail: darevalo@geomar.de

Yuri Artioli is a senior researcher at the Plymouth Marine Laboratory. His research interests include modelling impact of climate change and ocean acidification on marine biogeochemistry and ecosystems and their feedback on climate. Address: Plymouth Marine Laboratory, Prospect Place, The Hoe, Plymouth PL1 3DH, UK. e-mail: yuti@pml.ac.uk

Dawn M. Ashby is a senior Communications officer at Plymouth Marine Laboratory. Her interests include science communication and generation of non-academic impact from research. Address: Plymouth Marine Laboratory, Prospect Place, The Hoe, Plymouth PL1 3DH, UK. e-mail: daas@pml.ac.uk

Ian Brown is a marine Chemist at the Plymouth Marine Laboratory. His research interests are the ocean-atmosphere interaction of greenhouse gases in estuaries, coastal waters and the open ocean. Address: Plymouth Marine Laboratory, Prospect Place, The Hoe, Plymouth PL1 3DH, UK. e-mail: ib@pml.ac.uk

Hanna I. Campen is a doctoral researcher at the Chemical Oceanography Department of GEOMAR Helmholtz Centre for Ocean Research Kiel. Her research interests include the impact of global warming on marine biogeochemistry of the climate-relevant trace gases dimethyl sulphide and carbon monoxide in the Arctic Ocean. Address: GEOMAR Helmholtz-Zentrum Fu¨r Ozeanforschung Kiel, Chemische Ozeanographie, Du¨sternbrooker Weg 20, 24105 Kiel, Germany. e-mail: hcampen@geomar.de

Darren R. Clark is a senior research scientist at the Plymouth Marine Laboratory. His research interests include the biogeochemical cycling of carbon and nitrogen in marine systems. Address: Plymouth Marine Laboratory, Prospect Place, The Hoe, Plymouth PL1 3DH, UK. e-mail: drcl@pml.ac.uk

Vassilis Kitidis is a senior scientist at the Plymouth Marine Laboratory. His research interests in marine biogeochemistry include greenhouse gas cycling (CO2, CH4, N2O), carbonate chemistry, dissolved organic matter and nitrogen cycling from the ocean surface to the seabed. Address: Plymouth Marine Laboratory, Prospect Place, The Hoe, Plymouth PL1 3DH, UK. e-mail: vak@pml.ac.uk

Gennadi Lessin is a senior marine systems modeller at Plymouth Marine Laboratory. His research interests include development and application of modelling tools to study biogeochemical-ecological dynamics of the ocean and its response to natural and anthropogenic disturbance. His focus areas include benthic-pelagic interactions, dynamics of greenhouse gases and marine nitrogen cycle, as well as integrated analysis of ecosystem state and dynamics to inform management and policy. Address: Plymouth Marine Laboratory, Prospect Place, The Hoe, Plymouth PL1 3DH, UK. e-mail: gle@pml.ac.uk

Glen A. Tarran is a scientist at the Plymouth Marine Laboratory. His research interests include microbial community structure and dynamics and the role they play in the production and consumption of climatically-active gases. Address: Plymouth Marine Laboratory, Prospect Place, The Hoe, Plymouth PL1 3DH, UK. e-mail: gat@pml.ac.uk

Carol Turley is a senior scientist at the Plymouth Marine Laboratory. Her research interests include marine biogeochemistry and ocean policy at the international level. Address: Plymouth Marine Laboratory, Prospect Place, The Hoe, Plymouth PL1 3DH, UK. e-mail: ct@pml.ac.uk

Cite This Work

To export a reference to this article please select a referencing stye below:

Reference Copied to Clipboard.
Reference Copied to Clipboard.
Reference Copied to Clipboard.
Reference Copied to Clipboard.
Reference Copied to Clipboard.
Reference Copied to Clipboard.
Reference Copied to Clipboard.

Related Services

View all

Related Content

All Tags

Content relating to: "Environmental Science"

Environmental science is an interdisciplinary field focused on the study of the physical, chemical, and biological conditions of the environment and environmental effects on organisms, and solutions to environmental issues.

Related Articles

DMCA / Removal Request

If you are the original writer of this dissertation and no longer wish to have your work published on the UKDiss.com website then please: