Methane and nitrous oxide emissions from a subtropical coastal embayment (Moreton Bay, Australia)

https://doi.org/10.1016/j.jes.2014.06.049Get rights and content

Abstract

Surface water methane (CH4) and nitrous oxide (N2O) concentrations and fluxes were investigated in two subtropical coastal embayments (Bramble Bay and Deception Bay, which are part of the greater Moreton Bay, Australia). Measurements were done at 23 stations in seven campaigns covering different seasons during 2010–2012. Water–air fluxes were estimated using the Thin Boundary Layer approach with a combination of wind and currents-based models for the estimation of the gas transfer velocities. The two bays were strong sources of both CH4 and N2O with no significant differences in the degree of saturation of both gases between them during all measurement campaigns. Both CH4 and N2O concentrations had strong temporal but minimal spatial variability in both bays. During the seven seasons, CH4 varied between 500% and 4000% saturation while N2O varied between 128 and 255% in the two bays. Average seasonal CH4 fluxes for the two bays varied between 0.5 ± 0.2 and 6.0 ± 1.5 mg CH4/(m2·day) while N2O varied between 0.4 ± 0.1 and 1.6 ± 0.6 mg N2O/(m2·day). Weighted emissions (t CO2-e) were 63%–90% N2O dominated implying that a reduction in N2O inputs and/or nitrogen availability in the bays may significantly reduce the bays' greenhouse gas (GHG) budget. Emissions data for tropical and subtropical systems is still scarce. This work found subtropical bays to be significant aquatic sources of both CH4 and N2O and puts the estimated fluxes into the global context with measurements done from other climatic regions.

Introduction

Methane (CH4) and nitrous oxide (N2O) are two key potent greenhouse gases (GHGs). Both CH4 and N2O are long-lived atmospheric trace gases that remain chemically active for 8–12 years and 114–120 years, respectively (Ehhalt et al., 2001). They have respective global warming potentials of around 25 and 300 times that of carbon dioxide (IPCC, 2007, Myhre et al., 2013). N2O is also currently believed to be the single most important ozone depleting substance, a position it is likely to retain for years to come (Ravishankara et al., 2009). Of great concern is that both CH4 and N2O concentrations in the atmosphere have significantly increased since industrialisation and are still on the rise (IPCC, 1990, Stocker et al., 2013, Rigby et al., 2008). Yet the significance of a number of the drivers for this increase remains unclear and quantitatively the strengths of the various sources and sinks are still highly uncertain (Kirschke et al., 2013).

A significant portion of atmospheric and stratospheric CH4 and N2O is of biogenic origin and aquatic systems are believed to be significant sources (Bastviken et al., 2011, Bouwman et al., 1995, Khalil and Rasmussen, 1993, Nevison et al., 1995, Walter et al., 2007). Oceans are estimated to contribute 13%–25% of the global N2O emissions (Bouwman et al., 1995, Nevison et al., 1995) and 1%–4% of the global CH4 emissions (Cicerone and Oremland, 1988, Karl et al., 2008) with a significant proportion of the emissions in near shore areas likely influenced by discharges from rivers and estuaries (Borges and Abril, 2011). Great uncertainty, however, surrounds these estimates mainly due to scarcity of emission measurement data (Bange et al., 1994, Cicerone and Oremland, 1988, Nevison et al., 1995). Measurement data are particularly rare from tropical and subtropical aquatic systems as most marine and coastal system studies in these areas have mainly concentrated on upwelling regions (Charpentier et al., 2010, Codispoti et al., 1992, Naqvi et al., 2009, Naqvi et al., 2000). Application of emission factors from other climatic regions to these areas during regional and global GHG emissions budgeting is likely to cause uncertainty due to neglect of spatial variability, a well known occurrence even at a small areal extent (Bange et al., 1994, Bange et al., 1998, Zhang et al., 2006). Indeed both oceanic N2O and CH4 emissions are not uniformly distributed and it is likely that previous oceanic GHG budgets that were based on open ocean measurements significantly underestimated oceanic emissions (Bange, 2006). Biologically productive coastal and estuarine areas contribute around 60% of the oceanic N2O budget (Bange et al., 1996, Seitzinger et al., 2000) and up to 75% of the global oceanic CH4 budget (Bange et al., 1994). Besides biological processes, other causes for the high N2O and CH4 concentrations and fluxes in coastal areas may include effluent discharges from wastewater treatment plants (Hashimoto et al., 1999, Zhang et al., 2006), ground water input (LaMontagne et al., 2003, Ronen et al., 1988), input from connected creeks, rivers and estuaries (Ferrón et al., 2007, Musenze et al., 2014, Scranton and McShane, 1991, Zhang et al., 2006) and gas seeps (Amouroux et al., 2002, Reeburgh et al., 1991). Another challenge to regional and global GHG emission budgeting is that existing studies on which such estimates would be based do not adequately capture temporal variability as they are often of measurements done in a single season (Borges et al., 2011, Zhang et al., 2004). Consequently, the lack of information on temporal variability may constitute a large source of uncertainty during global emissions budgeting. There is, therefore, a need for more studies capturing both temporal and spatial variability from systems in diverse climatic divides to support adequate evaluation of regional and global N2O and CH4 budgets.

In this study we present a comprehensive assessment of CH4 and N2O concentrations and water–air emissions variability from two bays that are part of a subtropical coastal embayment in the southern hemisphere. CH4 and N2O concentration measurements were done at 23 stations in different seasons spanning over two years to adequately capture both spatial and temporal variability. Water–air fluxes were estimated using the Thin Boundary Layer Equation (TBLE) with a combination of wind and currents-based models for the estimation of the gas transfer velocity.

Section snippets

Physical setting and monitoring stations

Moreton Bay is 125 km long with an area of ca. 1500 km2 and a catchment area of ca. 22,000 km2. It has an average depth of 6.8 m and a tidal range of ca. 2 m (Dennison and Abal, 1999). It is lagoon–like in setting, with a series of offshore islands that control the movement of the pacific waters into and out of the bay (Fig. 1). It is associated with over 20 major rivers and creeks including the Brisbane River, which is the longest river in South East Queensland. Some of these rivers traverse

Physicochemical water parameters

The water column was always well mixed with only minor stratification during spring (Fig. 2a). The water column was also always well oxygenated but DO decreased from the surface towards the bottom in summer 2011 (Fig. 2b). Table 1 summarises other measured physicochemical water parameters. Salinity was remarkably low in summer 2011 compared to other seasons. Chlorophyll-a concentrations were closely similar in the two bays and peaked in summer 2011 (Fig. 3a). Total nitrogen (Fig. 3b), organic

Bays are a strong source of both N2O and CH4

Our study found the bays supersaturated with both N2O and CH4 in all seasons during which measurements were done. Data from tropical and subtropical systems is still scarce (Naqvi et al., 2009) but Oudot et al. (1990) reported the degree of surface water N2O saturation of 123%–132% in the tropical Atlantic Ocean close to the West African coast of Senegal. These tropical measurements are on the lower tail end of our observed degree of N2O saturation (128%–255%), which is also generally higher

Acknowledgments

This work was funded by the Australian Research Council (ARC), Healthy Waterways Ltd and Seqwater through an industry linkage grant (ARC Linkage project # LP100100325). We thank Dr. Alberto Vieira Borges, Dr. Philip M. Nyenje, Dr. R. Kulabako, Dr. U. Bagampadde and Irene Nansubuga for the discussions during data analysis. The authors greatly appreciate the support of EHMP–DERM management, tech staff and crews of Seratta and Seriops cruises (2010–2012) including but not limited to: M. Holmes, R.

References (78)

  • M.G. LaMontagne et al.

    Nitrous oxide sources and sinks in coastal aquifers and coupled estuarine receiving waters

    Sci. Total Environ.

    (2003)
  • R.S. Musenze et al.

    Methane and nitrous oxide emissions from a subtropical estuary (the Brisbane River estuary, Australia)

    Sci. Total Environ.

    (2014)
  • S.W.A. Naqvi et al.

    Nitrous oxide in the western Bay of Bengal

    Mar. Chem.

    (1994)
  • C. Oudot et al.

    Nitrous oxide production in the tropical Atlantic Ocean

    Deep Sea Res. Part A. Oceanogr. Res. Papers

    (1990)
  • C. Oudot et al.

    Transatlantic equatorial distribution of nitrous oxide and methane

    Deep Sea Res. Part I: Oceanogr. Res. Papers

    (2002)
  • W.S. Reeburgh et al.

    Black Sea methane geochemistry

    Deep Sea Res. Part A. Oceanogr. Res. Papers

    (1991)
  • M.I. Scranton et al.

    Methane fluxes in the southern North Sea: the role of European rivers

    Cont. Shelf Res.

    (1991)
  • S.P. Seitzinger et al.

    Global distribution of N2O emissions from aquatic systems: natural emissions and anthropogenic effects

    Chemosphere Global Change Sci.

    (2000)
  • R.C. Upstill-Goddard

    Air-sea gas exchange in the coastal zone

    Estuar. Coast. Shelf Sci.

    (2006)
  • T. Usui et al.

    N2O production, nitrification and denitrification in an estuarine sediment

    Estuar. Coast. Shelf Sci.

    (2001)
  • R.F. Weiss et al.

    Nitrous oxide solubility in water and seawater

    Mar. Chem.

    (1980)
  • M.J. Whiticar et al.

    Methane oxidation in sediment and water column environments—isotope evidence

    Org. Geochem.

    (1986)
  • R.J. Wilcock

    Methyl chloride as a gas-tracer for measuring stream reaeration coefficients—II: stream studies

    Water Res.

    (1984)
  • Y.P. Xing et al.

    Methane and carbon dioxide fluxes from a shallow hypereutrophic subtropical Lake in China

    Atmos. Environ.

    (2005)
  • G.L. Zhang et al.

    Distributions, sources and atmospheric fluxes of nitrous oxide in Jiaozhou Bay

    Estuar. Coast. Shelf Sci.

    (2006)
  • APHA

    Standard Methods for the Examination of Water and Wastewater

    (1998)
  • H.W. Bange et al.

    Methane in the Baltic and North Seas and a reassessment of the marine emissions of methane

    Global Biogeochem. Cycles

    (1994)
  • H.W. Bange et al.

    Nitrous oxide in coastal waters

    Global Biogeochem. Cycles

    (1996)
  • J. Barnes et al.

    N2O seasonal distributions and air–sea exchange in UK estuaries: implications for the tropospheric N2O source from European coastal waters

    J. Geophys. Res. Biogeosci.

    (2011)
  • D. Bastviken et al.

    Freshwater methane emissions offset the continental carbon sink

    Science

    (2011)
  • A.V. Borges et al.

    Variability of the gas transfer velocity of CO2 in a macrotidal estuary (the Scheldt)

    Estuaries

    (2004)
  • A.F. Bouwman et al.

    Uncertainties in the global source distribution of nitrous oxide

    J. Geophys. Res.

    (1995)
  • J. Charpentier et al.

    Nitrous oxide fluxes in the central and eastern South Pacific

    Glob. Biogeochem. Cycle

    (2010)
  • C.R. Chu et al.

    Wind and stream flow induced reaeration

    J. Environ. Eng.

    (2003)
  • R.J. Cicerone et al.

    Biogeochemical aspects of atmospheric methane

    Global Biogeochem. Cycles

    (1988)
  • J.F. Clark et al.

    Gas exchange rates in the tidal Hudson river using a dual tracer technique

    Tellus B

    (1994)
  • L.A. Codispoti et al.

    On the nitrous oxide flux from productive regions that contain low oxygen waters

  • CSIRO

    South East Queensland Receiving Water Quality Model V3 (RWQM3)

    (2010)
  • F.J. Cynar et al.

    The distribution of methane in the upper waters of the Southern California Bight

    J. Geophys. Res. C Oceans

    (1992)
  • Cited by (15)

    • Inhibition effect of magnetic field on nitrous oxide emission from sequencing batch reactor treating domestic wastewater at low temperature

      2020, Journal of Environmental Sciences (China)
      Citation Excerpt :

      During wastewater treatment, considerable N2O can be produced (Liu et al., 2016), especially in the denitrification stage of the nitrogen removal process (Wu et al., 2013). Greenhouse gases (GHG) were emitted, namely N2O, CO2 and CH4 (Musenze et al., 2015), of which N2O accounted for 26% (Kampschreur et al., 2009). N2O emission accounted for 0–22% of the removed nitrogen in wastewater treatment adopting biological nitrogen removal (BNR) processes (Peng et al., 2015).

    • Seasonal variations of nitrous oxide fluxes and soil denitrification rates in subtropical freshwater and brackish tidal marshes of the Min River estuary

      2018, Science of the Total Environment
      Citation Excerpt :

      Temperature would also affect the physical properties of soil, as well as the solubility of N2O within both the pore water and the overlying water (Weiss and Price, 1980; Wanninkhof, 1992). Temperature would also affect the water-air gas transfer rate (Guo et al., 2013; Musenze et al., 2014, 2015; Zhu et al., 2014). The significant positive correlation between N2O flux and temperature (Table S2) indicates that temperature is an important control of the seasonal variability of N2O flux.

    • Spartina alterniflora alters ecosystem DMS and CH<inf>4</inf> emissions and their relationship along interacting tidal and vegetation gradients within a coastal salt marsh in Eastern China

      2017, Atmospheric Environment
      Citation Excerpt :

      Coastal marshes have consistently been confirmed as significant sources of DMS and CH4 (Kiene and Visscher, 1987; Kiene, 1988; De Souza and Yoch, 1996; DeLaune et al., 2002; Yuan et al., 2014). Previous studies have found methanogenesis to occur when CO2, methyl compounds, or acetate are available as substrates (Oremland and Polcin, 1982; King et al., 1983; Oremland et al., 1991; Whiticar, 1999; Parkes et al., 2012; Penger et al., 2012; Costa and Leigh, 2014; Toyoda et al., 2014; Li et al., 2015b; Lin et al., 2015; Maltby et al., 2015; Musenze et al., 2015; Sorokin et al., 2015; Sun et al., 2015). DMS can also be a non-competitive substrate for methanogenesis (Zinder and Brock, 1978; Jørgensen, 1982; Kiene et al., 1986; Kiene and Visscher, 1987; Fu and Metcalf, 2015), which suggests that production and emission of DMS and CH4 are likely to be positively correlated.

    View all citing articles on Scopus
    1

    Department of Civil and Environmental Engineering, College of Engineering, Design, Arts and Technology, Makerere University, P.O. Box 7062, Kampala, Uganda.

    View full text