Journal list menu

Volume 61, Issue S1 p. S267-S282
Article
Free Access

Iron-dependent anaerobic oxidation of methane in coastal surface sediments: Potential controls and impact

Jurjen Rooze

Corresponding Author

Jurjen Rooze

Department of Earth Sciences, Faculty of Geosciences, Utrecht University, PO Box 80021, 3508 TA the Netherlands

Correspondence: [email protected]Search for more papers by this author
Matthias Egger

Matthias Egger

Department of Earth Sciences, Faculty of Geosciences, Utrecht University, PO Box 80021, 3508 TA the Netherlands

Search for more papers by this author
Iana Tsandev

Iana Tsandev

Department of Earth Sciences, Faculty of Geosciences, Utrecht University, PO Box 80021, 3508 TA the Netherlands

Search for more papers by this author
Caroline P. Slomp

Caroline P. Slomp

Department of Earth Sciences, Faculty of Geosciences, Utrecht University, PO Box 80021, 3508 TA the Netherlands

Search for more papers by this author
First published: 07 April 2016
Citations: 50

Special Issue: Methane Emissions from Oceans, Wetlands, and Freshwater Habitats: New Perspectives and Feedbacks on Climate Edited by: Kimberly Wickland and Leila Hamdan

Abstract

Anaerobic oxidation of methane (AOM) is an important process of methane (CH4) removal in sediments. Various studies suggest that AOM coupled to iron oxide (Fe(OH)3) reduction (Fe-AOM) may complement sulfate-driven AOM in CH4-rich sediments. Here, we apply a transient reaction-transport model to depth profiles of key porewater and sediment constituents for a site in the Bothnian Sea where Fe-AOM has been suggested to occur. At the site, increased eutrophication has led to an upward shift of the sulfate-methane transition zone, submerging Fe(OH)3 in a zone with high CH4 concentrations. Fe-AOM is thought to lead to a strong accumulation of dissolved iron (Fe2+) in the porewater. Results of a sensitivity analysis identify three potential controls on the occurrence of Fe-AOM in coastal surface sediments: (1) bottom-water sulfate ( urn:x-wiley:00243590:media:lno10275:lno10275-math-0001) concentrations, (2) Fe(OH)3 availability, and (3) organic matter (OM) loading. In-situ CH4 production is particularly sensitive to the OM loading and urn:x-wiley:00243590:media:lno10275:lno10275-math-0002 bottom-water concentration, with higher urn:x-wiley:00243590:media:lno10275:lno10275-math-0003 concentrations significantly inhibiting methanogenesis and decreasing the potential rates of Fe-AOM. We find that only environments with a low salinity and a relatively high Fe(OH)3 loading allow for Fe-AOM to occur in surface sediments. This suggests that Fe-AOM in surface sediments is restricted to areas with relatively high rates of sediment deposition such as estuaries and other nearshore systems. By enhancing porewater Fe2+ concentrations in surface sediments and the flux of Fe2+ from sediments to the overlying water, Fe-AOM may contribute to the lateral transfer of iron (“iron shuttling”) from the coastal zone to deep basins.

Methane (CH4) is produced in sediments by methanogens during the last step of the microbial breakdown of organic matter (OM). As CH4 is a 28–34 times more potent greenhouse gas than carbon dioxide (CO2) on a centennial time scale (Myhre et al. 2013), understanding its fate in sediments is important. The oxidation of CH4 to CO2 is facilitated by methanotrophs, which gain energy from these reactions. Various electron acceptors can be involved in this process. Anaerobic oxidation of CH4 coupled to sulfate ( urn:x-wiley:00243590:media:lno10275:lno10275-math-0004) reduction (SO4-AOM) is considered to be the largest CH4 sink in marine sediments (Reeburgh 2007; Knittel and Boetius 2009). In the upper few millimeters of the sediment, CH4 oxidation can also be coupled to oxygen (O2) reduction (Martens and Berner 1974; Boetius et al. 2000; Reeburgh 2007). Laboratory studies have shown that under thermodynamically favorable conditions nitrite and nitrate may also be used to oxidize CH4 (Raghoebarsing et al. 2006; Ettwig et al. 2008; Segarra et al. 2013).

Metal oxides, specifically iron oxides (FeOH3) and manganese oxides (MnO2), are another group of electron acceptors posited to facilitate anaerobic oxidation of methane (AOM) in the absence of urn:x-wiley:00243590:media:lno10275:lno10275-math-0005. The first evidence for a potential coupling between AOM and metal oxide reduction in marine CH4-seep sediments was deduced from incubation experiments with 13C-labeled CH4 (Beal et al. 2009). When MnO2 or Fe(OH)3 was added, 13CO2 accumulated over time, whereas this did not occur in the absence of other electron acceptors or if the samples were autoclaved. The net reaction (Eq. 1) proposed for AOM coupled to Fe(OH)3 reduction (Fe-AOM) by Beal et al. (2009) consumes eight moles of Fe(OH)3 for each mole of CH4 oxidized:
urn:x-wiley:00243590:media:lno10275:lno10275-math-0006(1)

Other incubation experiments using brackish sediments (Segarra et al. 2013; Egger et al. 2015a) also point to a decoupling of AOM from urn:x-wiley:00243590:media:lno10275:lno10275-math-0007 reduction and the potential role of Fe(OH)3 as the electron acceptor.

In-situ observation of this process in natural sediments has been more challenging, as a clear microbiological or geochemical signature of Fe-AOM has not yet been established. Environments with the potential to host the Fe-AOM reaction have been identified in numerous studies and include lake sediments (Adler et al. 2011; Sivan et al. 2011; Norði et al. 2013), hydrothermal vent sediments (Wankel et al. 2012), brackish surface sediments (Segarra et al. 2013; Slomp et al. 2013; Egger et al. 2015a) and the deep subsurface of marine sediments (Riedinger et al. 2014; Treude et al. 2014). Some of these studies (Slomp et al. 2013; Riedinger et al. 2014; Egger et al. 2015a) argue for the presence of AOM coupled to the reduction of metal oxides on the basis of strong circumstantial geochemical evidence. However, the microbes facilitating these reactions have yet to be identified. Furthermore, laboratory experiments with seep sediments recently revealed that the presence of Fe(OH)3 may stimulate SO4-AOM, likely through a cryptic sulfur cycle (Holmkvist et al. 2011; Sivan et al. 2014). Fe-AOM could thus also be, at least partly, the result of indirect iron stimulated SO4-AOM.

Little is known about the controls on Fe-AOM rates and what environmental conditions may favor this reaction. The concurrent presence of dissolved CH4 and abundant reducible Fe(OH)3 in marine sediments requires burial of Fe(OH)3 below the zone of urn:x-wiley:00243590:media:lno10275:lno10275-math-0008 reduction. Previous studies suggest that such a situation can occur in sediments that either receive a high input of Fe(OH)3 or are subject to transient diagenesis (Riedinger et al. 2014; Egger et al. 2015a). In coastal sediments, increased inputs of OM to the sediment linked to anthropogenic eutrophication can induce such transient diagenesis and can lead to a vertical upward migration of the urn:x-wiley:00243590:media:lno10275:lno10275-math-0009/CH4 transition zone (SMTZ) and the presence of Fe(OH)3 below the SMTZ (Slomp et al. 2013; Egger et al. 2015a). To date, the biogeochemical significance of the Fe-AOM reaction in coastal sediments remains unclear.

An important consideration regarding the occurrence of Fe-AOM in near surface sediments is whether the reaction enhances the release of dissolved iron (Fe2+) from the sediment to the overlying water column. In the case of the Bothnian Sea, Fe2+ may be transported to the Baltic Proper, thus potentially contributing to the shelf-to-basin shuttling of iron (Fehr et al. 2010; Jilbert and Slomp 2013). Following this mechanism, Fe2+ released from oxic or hypoxic shelf sediments is transported toward adjacent deeper basins, either in the form of Fe2+ (in an anoxic water-column) or as reactive colloidal or nanoparticulate Fe(OH)3 (in oxic waters). If the basins are anoxic and sulfidic (euxinic), iron is trapped in the form of iron sulfides (FeS and FeS2), and is subsequently permanently buried in the sediment (Wijsman et al. 2001; Lyons and Severmann 2006; Raiswell and Canfield 2012). Some of the laterally transported iron may also reach the surface waters where it may contribute to the development of cyanobacterial blooms (Breitbarth et al. 2009). This highlights how enhanced Fe2+ effluxes due to Fe-AOM could potentially have significant implications for the functioning of coastal ecosystems.

In this study, we develop a reactive transport model to simulate the cycling of carbon, nitrogen, oxygen, phosphorus, sulfur, and iron in coastal marine sediments subject to anthropogenic eutrophication. We use our model to simulate selected porewater and sediment profiles from the Bothnian Sea as published by Egger et al. (2015a, 2015b). We then perform a sensitivity study to assess which factors could control the rate of Fe-AOM in coastal marine sediments with the additional aim of identifying environments where Fe-AOM is likely to take place. We also look for potential geochemical signatures—in the form of a profile or a combination of profiles—that may signal the occurrence of Fe-AOM (e.g., high dissolved Fe2+ below the SMTZ). We find that a relatively low water column salinity (and correspondingly low urn:x-wiley:00243590:media:lno10275:lno10275-math-0010 concentration) and high Fe(OH)3 availability to be critical parameters controlling Fe-AOM.

Methods

Model

A transient diagenetic model was developed simulating the cycling of 20 chemical species in the sediment (Table 1). The model domain is represented by a one-dimensional grid of 550 cells that captures the interval from the sediment-water interface (SWI) to a depth of 55 cm. Key environmental parameters, such as the porosity, temperature and transport characteristics of the sediment were either taken from literature or constrained from the model (Table 2). The model accounts for biogeochemical reactions (Table 3) and transport processes through burial and bioturbation, as well as molecular/ionic diffusion for dissolved species (Soetaert et al. 1996; Wang and Van Cappellen 1996; Boudreau 1997). The distribution of solid and dissolved species is described by Eqs. 2 and 3, respectively:
urn:x-wiley:00243590:media:lno10275:lno10275-math-0011(2)
urn:x-wiley:00243590:media:lno10275:lno10275-math-0012(3)
where Ci and Cj denote concentrations of solid and aqueous species, respectively. ϕ represents the sediment porosity, Db the bioturbation coefficient, Dj the diffusion coefficients of dissolved species in the sediment, and v and u the advective velocities of the solid and the aqueous phases, respectively (Table 2). The R term denotes the net source or sink of a chemical species from the sum of chemical reactions it participates in.
Table 1. Chemical species included in the model
Species Notation Type
Organic mattera urn:x-wiley:00243590:media:lno10275:lno10275-math-0013 Solid
Oxygen O2 Solute
Nitrate urn:x-wiley:00243590:media:lno10275:lno10275-math-0014 Solute
Sulfate urn:x-wiley:00243590:media:lno10275:lno10275-math-0015 Solute
Iron oxideb Fe urn:x-wiley:00243590:media:lno10275:lno10275-math-0016 Solid
Methane CH4 Solute
Iron Fe2+ Solute
Ammonium urn:x-wiley:00243590:media:lno10275:lno10275-math-0017 Solute
Hydrogen sulfidec ΣH2S Solute
Elemental sulfur S0 Solid
Iron monosulfide FeS Solid
Pyrite FeS2 Solid
Phosphate urn:x-wiley:00243590:media:lno10275:lno10275-math-0018 Solute
Dissolved inorganic carbon DIC Solute
Iron-bound phosphorusd FeP Solid
Siderite FeCO3 Solid
  • a Superscript “α,” “β,” and “γ” denote highly reactive, less reactive, and refractory organic matter, respectively.
  • b Superscript “α,” “β,” and “γ” denote an increase in the degree of crystallinity.
  • c ΣH2S includes H2S, HS, and S2−.
  • d Iron-bound P accounts for both Fe(OH)3-bound P (FePox) and P in vivianite (FePviv), i.e., FeP = FePox + FePviv.
To account for sediment compaction, the porosity is described by
urn:x-wiley:00243590:media:lno10275:lno10275-math-0019(4)
where x is the depth in the sediment, ϕ0 the porosity at the top of the sediment, ϕ the porosity at depth in the sediment, and γ the porosity attenuation factor (e-folding distance, see Table 2).
Table 2. Environmental parameters used in the model
Description Symbol Value or expression Source
Porosity at surface ϕ0 0.94 b
Porosity at depth ϕ 0.89 b
Porosity e-folding distance γ 5 cm a
Sediment density ρ 2.65 g cm−3 c
Temperature T 4 °C b
Salinity S 6 b
Advective velocity (cm y−1) of solids at depth v urn:x-wiley:00243590:media:lno10275:lno10275-math-0020
Bioturbation coefficient at SWI Db0 2.7 cm2 y−1 c
Mixed layer depth ζ 4 cm a
  • Sources: (a) Model constrained; (b) Egger et al. (2015a); (c) Reed et al. (2011). Note: At this site, lugworms were absent (Norkko, personal communication), making bioirrigation a negligible process.
Diffusion coefficients Dj in the sediment were calculated as a function of temperature and salinity (Table 2) following Boudreau (1997) and include a correction for the tortuosity θ in the porous medium (Boudreau 1996), where θ is calculated as follows
urn:x-wiley:00243590:media:lno10275:lno10275-math-0021(5)
The coefficient Db was used to account for transport of solids and solutes caused by mixing of the sediment through macrofaunal activity, i.e., bioturbation, and was implemented as
urn:x-wiley:00243590:media:lno10275:lno10275-math-0022(6)
where Db0 denotes the bioturbation coefficient near the SWI and ζ the mixed layer depth (Table 2). Following Meysman et al. (2005), the advective velocities v and u of solids and solutes, respectively, caused by sediment accumulation were described by depth-dependent functions to account for changes in the porosity.

The reaction network in the model describes the breakdown of OM, the reoxidation of reduced metabolites, and various mineral formation and dissolution reactions (Table 3). OM degradation occurs through respiratory reactions in which O2, Fe(OH)3, urn:x-wiley:00243590:media:lno10275:lno10275-math-0023, and urn:x-wiley:00243590:media:lno10275:lno10275-math-0024 serve as terminal electron acceptors and through methanogenesis. Following the multi-G approach (Jørgensen 1978; Westrich and Berner 1984), three state variables for OM were used to distinguish between highly reactive (α), less reactive (β), and nonreactive (γ). OM includes organic carbon (C), organic nitrogen (N), and organic phosphorus (P) in a C : N : P ratio of 106 : 15.4 : 1. For Fe(OH)3 an α, β and γ phase were included to denote variations in crystallinity, which affect the reactivity of Fe(OH)3 toward OM and ΣH2S (Tables 3 and 4). The α phase reacts with both OM and ΣH2S, the β phase persists in the sediment past OM degradation but is susceptible to sulfide scavenging, while the γ phase is not reactive at all. The segregated reactivity of the different Fe(OH)3 forms was necessary to allow Fe(OH)3 to persist past the zone of organoclastic Fe(OH)3 reduction (avoiding reaction with OM) and to be reduced at further depth creating the observed Fe2+ accumulation. The model includes oxidation of CH4 coupled to O2, urn:x-wiley:00243590:media:lno10275:lno10275-math-0025, and Fe(OH)3, where CH4 is assumed to react with both the α and β phase of Fe(OH)3. Although MnO2 has also been shown to be a thermodynamically favorable electron acceptor for AOM (Beal et al. 2009), this study focuses on the potential role of Fe(OH)3 for AOM only due to the relatively low sedimentary manganese content below the SMTZ at site US5B ( < 40 µmol/µmolg−1) when compared to iron (Egger et al. 2015b) and the likely presence of most of the manganese in the form of manganese carbonate at the relevant depths.

Table 3. Reaction pathways and stoichiometries implemented in the model
Primary redox reactionsa
OMα,β + a O2a CO2 + b urn:x-wiley:00243590:media:lno10275:lno10275-math-0026 + c urn:x-wiley:00243590:media:lno10275:lno10275-math-0027 + a H2O R1
OMα,β + 0.8a urn:x-wiley:00243590:media:lno10275:lno10275-math-0028a CO2 + b urn:x-wiley:00243590:media:lno10275:lno10275-math-0029 + c urn:x-wiley:00243590:media:lno10275:lno10275-math-0030 + 0.6a H2O + 0.4a N2 + 0.8a OH R2
OMα,β + 4a urn:x-wiley:00243590:media:lno10275:lno10275-math-0031 + 4α FePoxa CO2 + b urn:x-wiley:00243590:media:lno10275:lno10275-math-0032 + (c + 4α) urn:x-wiley:00243590:media:lno10275:lno10275-math-0033 + 3a H2O + 4a Fe2+ + 8a OH R3
OMα,β + 0.5a urn:x-wiley:00243590:media:lno10275:lno10275-math-0034a CO2 + b urn:x-wiley:00243590:media:lno10275:lno10275-math-0035 + c urn:x-wiley:00243590:media:lno10275:lno10275-math-0036 + 0.5a H2S + a OH R4
OMα,β→ 0.5a CO2 + b urn:x-wiley:00243590:media:lno10275:lno10275-math-0037 + c urn:x-wiley:00243590:media:lno10275:lno10275-math-0038 + 0.5a CH4 R5
Other reactions
2 O2 +  urn:x-wiley:00243590:media:lno10275:lno10275-math-0039 + 2 urn:x-wiley:00243590:media:lno10275:lno10275-math-0040urn:x-wiley:00243590:media:lno10275:lno10275-math-0041 + 2 CO2 + 3 H2O R6
O2 + 4 Fe2+ + 8 urn:x-wiley:00243590:media:lno10275:lno10275-math-0042 + 2 H2O + 4χα urn:x-wiley:00243590:media:lno10275:lno10275-math-0043→ 4 urn:x-wiley:00243590:media:lno10275:lno10275-math-0044 + 4χα FePox + 8 CO2 R7
2 O2 + FeS→ urn:x-wiley:00243590:media:lno10275:lno10275-math-0045 + Fe2+ R8
7 O2 + 2 FeS2 + 4 OH→ 4 urn:x-wiley:00243590:media:lno10275:lno10275-math-0046 + 2 Fe2+ + 2 H2O R9
2 O2 + H2S + 2 urn:x-wiley:00243590:media:lno10275:lno10275-math-0047urn:x-wiley:00243590:media:lno10275:lno10275-math-0048 + 2 CO2 + 2 H2O R10
2 O2 + CH4→ CO2 + 2 H2O R11
2 urn:x-wiley:00243590:media:lno10275:lno10275-math-0049 + 2χα,β FeP + H2S + 4 CO2→ 2 Fe2+ + 2χα,β urn:x-wiley:00243590:media:lno10275:lno10275-math-0050 + S0 + 4 urn:x-wiley:00243590:media:lno10275:lno10275-math-0051 + 2 H2O R12
Fe2+ + H2S + 2 urn:x-wiley:00243590:media:lno10275:lno10275-math-0052→ FeS + 2 CO2 + 2 H2O R13
urn:x-wiley:00243590:media:lno10275:lno10275-math-0053 + CH4 + CO2→ 2 urn:x-wiley:00243590:media:lno10275:lno10275-math-0054 + H2S R14
CH4 + 8 urn:x-wiley:00243590:media:lno10275:lno10275-math-0055 + 8χα,β FePoxurn:x-wiley:00243590:media:lno10275:lno10275-math-0056 + 8 Fe2+ + 8χα,β urn:x-wiley:00243590:media:lno10275:lno10275-math-0057 + 6 H2O + 15 OH R15
4 S0 + 2 H2O + 2 OH→ 3 H2S +  urn:x-wiley:00243590:media:lno10275:lno10275-math-0058 R16
FeS + S0→ FeS2 R17
urn:x-wiley:00243590:media:lno10275:lno10275-math-0059 + (χαχβ) FePox→ Fe urn:x-wiley:00243590:media:lno10275:lno10275-math-0060 + (χαχβ) urn:x-wiley:00243590:media:lno10275:lno10275-math-0061 R18
3 Fe2+ + 2 urn:x-wiley:00243590:media:lno10275:lno10275-math-0062→ 2 FePviv R19
Fe2+ +  urn:x-wiley:00243590:media:lno10275:lno10275-math-0063 + OH→ FeCO3 + H2O R20
FeS + H2S→ FeS2 + H2 R21
urn:x-wiley:00243590:media:lno10275:lno10275-math-0064 + (χα,βmaxχα,β) urn:x-wiley:00243590:media:lno10275:lno10275-math-0065urn:x-wiley:00243590:media:lno10275:lno10275-math-0066 + FePox R22
  • a Organic matter (OM) is of the form (CH2O)a( urn:x-wiley:00243590:media:lno10275:lno10275-math-0067)b( urn:x-wiley:00243590:media:lno10275:lno10275-math-0068)c. The “a,” “b,” and “c” refer to the C : N : P ratio in organic matter. χα,β,γ denote the P : Fe ratio of urn:x-wiley:00243590:media:lno10275:lno10275-math-0069, which were estimated based on the sedimentary profiles of urn:x-wiley:00243590:media:lno10275:lno10275-math-0070 and FePox (χα = 0.18, χβ = 0.1 and χγ = 0.06).
Table 4. Reaction equations implemented in the model
urn:x-wiley:00243590:media:lno10275:lno10275-math-0071
urn:x-wiley:00243590:media:lno10275:lno10275-math-0072
urn:x-wiley:00243590:media:lno10275:lno10275-math-0073
urn:x-wiley:00243590:media:lno10275:lno10275-math-0074
urn:x-wiley:00243590:media:lno10275:lno10275-math-0075
urn:x-wiley:00243590:media:lno10275:lno10275-math-0076
urn:x-wiley:00243590:media:lno10275:lno10275-math-0077
urn:x-wiley:00243590:media:lno10275:lno10275-math-0078
urn:x-wiley:00243590:media:lno10275:lno10275-math-0079
urn:x-wiley:00243590:media:lno10275:lno10275-math-0080
urn:x-wiley:00243590:media:lno10275:lno10275-math-0081
urn:x-wiley:00243590:media:lno10275:lno10275-math-0082
urn:x-wiley:00243590:media:lno10275:lno10275-math-0083
urn:x-wiley:00243590:media:lno10275:lno10275-math-0084
urn:x-wiley:00243590:media:lno10275:lno10275-math-0085
urn:x-wiley:00243590:media:lno10275:lno10275-math-0086
urn:x-wiley:00243590:media:lno10275:lno10275-math-0087
urn:x-wiley:00243590:media:lno10275:lno10275-math-0088
urn:x-wiley:00243590:media:lno10275:lno10275-math-0089
urn:x-wiley:00243590:media:lno10275:lno10275-math-0090
urn:x-wiley:00243590:media:lno10275:lno10275-math-0091
urn:x-wiley:00243590:media:lno10275:lno10275-math-0092
  • The superscript i denotes the highly reactive α and less reactive β fractions of organic matter; superscript j denotes the less crystalline α and more crystalline β fractions of iron oxyhydroxides.

Dissolved inorganic carbon (DIC) and porewater phosphate ( urn:x-wiley:00243590:media:lno10275:lno10275-math-0093) were included in the model to allow formation of the reduced Fe minerals siderite and vivianite, which act as key sinks for porewater Fe2+ (Tables 3 and 4). DIC was calculated as the sum of the carbon in CO2 and urn:x-wiley:00243590:media:lno10275:lno10275-math-0094 that was produced or consumed in the modeled reactions (Table 3). Dissolved urn:x-wiley:00243590:media:lno10275:lno10275-math-0095 was allowed to adsorb to existing urn:x-wiley:00243590:media:lno10275:lno10275-math-0096 (see R22, Table 3). It was assumed that the initial P:Fe ratios (χ of urn:x-wiley:00243590:media:lno10275:lno10275-math-0097 arriving at the sediment surface (χα = 0.18, χβ = 0.1) were lower than an estimated maximum binding capacity of urn:x-wiley:00243590:media:lno10275:lno10275-math-0098 under high urn:x-wiley:00243590:media:lno10275:lno10275-math-0099 porewater concentrations ( urn:x-wiley:00243590:media:lno10275:lno10275-math-0100 = 0.25, urn:x-wiley:00243590:media:lno10275:lno10275-math-0101 = 0.14). The boundary conditions at the SWI were specified as variable, time-dependent fluxes for total sediment accumulation and inputs of OM and Fe(OH)3 (see model application section), and as constant fluxes for all other solids and fixed concentrations for dissolved species (Table 5). For all species a zero gradient boundary condition was set at the bottom of the model domain. Bottom-water O2 and urn:x-wiley:00243590:media:lno10275:lno10275-math-0102 concentrations were imposed, while the concentrations of reduced species were set to zero. All CH4 was assumed to be produced within the model domain.

Table 5. Boundary conditions in the model. See Fig. 1 for the time-dependent OM and Fe(OH)3 fluxes at the SWI. Flux of OM in units of C. The flux of FeCO3 was adjusted to the time-dependent sedimentation rate to allow for a constant background concentration of FeCO3 of 120 µmol g−1 with depth in the sediment
Solids Flux at SWI
urn:x-wiley:00243590:media:lno10275:lno10275-math-0103 0 mol m−2 y−1
urn:x-wiley:00243590:media:lno10275:lno10275-math-0104 0 mol m−2 y−1
urn:x-wiley:00243590:media:lno10275:lno10275-math-0105 0 mol m−2 y−1
urn:x-wiley:00243590:media:lno10275:lno10275-math-0106 0.1 mol m−2 y−1
urn:x-wiley:00243590:media:lno10275:lno10275-math-0107 0.02–0.76 mol m−2 y−1
Solutes Bottom-water concentration
urn:x-wiley:00243590:media:lno10275:lno10275-math-0108 0.18 mol m−3
urn:x-wiley:00243590:media:lno10275:lno10275-math-0109 10 × 10−3 mol m−3
urn:x-wiley:00243590:media:lno10275:lno10275-math-0110 4.8 mol m−3
urn:x-wiley:00243590:media:lno10275:lno10275-math-0111 0 mol m−3
urn:x-wiley:00243590:media:lno10275:lno10275-math-0112 0 mol m−3
urn:x-wiley:00243590:media:lno10275:lno10275-math-0113 0 mol m−3
urn:x-wiley:00243590:media:lno10275:lno10275-math-0114 10 × 10−3 mol m−3
CDIC 0.15 mol m−3
  • For all chemical species a zero-gradient boundary condition was specified at the bottom of the model domain.

The model code was written in R using the marelac (Soetaert et al. 2010) geochemical dataset package and the ReacTran (Soetaert and Meysman 2012) package to calculate the transport in porous media. The ordinary differential equations were solved with the lsoda integrator algorithm (Hindmarsh 1983; Petzold 1983).

Model application and sensitivity analyses

The model was applied to a site in the Bothnian Sea, which is a brackish coastal basin with an average bottom water salinity of 5–6. The site considered, US5B, is located in the deepest part of the basin at 214 m water depth and is characterized by fine-grained, organic-rich sediments. The recent upward displacement of the SMTZ at this site and subsequent fixation of its position near the sediment surface reflects strong temporal changes in OM input over the past decades that have been attributed to a period of eutrophication followed by a recovery phase (Slomp et al. 2013; Egger et al. 2015a, 2015b). These observations are in accordance with long-term monitoring studies based on inorganic nutrients, water transparency and chlorophyll a in the Bothnian Sea that reveal substantial changes in anthropogenic loading of land-derived nutrients over recent decades (Fleming-Lehtinen et al. 2008; Savchuk et al. 2008; Fleming-Lehtinen and Laamanen 2012). Sedimentation rates at this site have previously been estimated to be high (up to 1.9 cm y−1) and have been shown to vary in space and time (Mattila et al. 2006; Egger et al. 2015b). Repeated sampling campaigns indicate that the depth of the peak in solid phase sulfur associated with the SMTZ did not change position relative to the sediment-water interface between 2008 and 2012, suggesting a recent decline in sedimentation rate (Egger et al. 2015b).

The model was calibrated to depth profiles of sediment organic carbon, porosity, solid sulfur (including FeS and FeS2), Fe(OH)3, siderite, iron-bound phosphorus (P) (which represents both P associated with Fe(OH)3 (FePox) and P in vivianite (FePviv)) and porewater profiles of CH4, urn:x-wiley:00243590:media:lno10275:lno10275-math-0115, urn:x-wiley:00243590:media:lno10275:lno10275-math-0116, ΣH2S, Fe2+, and urn:x-wiley:00243590:media:lno10275:lno10275-math-0117 that were collected at site US5B in August 2012 (Egger et al. 2015a, 2015b). Reaction parameters were mostly taken from the literature (Table 6). Where parameters were not available or no fit to the data could be obtained with existing parameter ranges, reaction parameters were constrained using the model. The relevant parameters are urn:x-wiley:00243590:media:lno10275:lno10275-math-0118, k8, k10, k12, k14, k15, k16, and k17 (Table 6) and thus include the rate constant of AOM coupled to Fe(OH)3 (k10). Rate constants for the reaction of ΣH2S with urn:x-wiley:00243590:media:lno10275:lno10275-math-0119 and urn:x-wiley:00243590:media:lno10275:lno10275-math-0120 differed slightly ( urn:x-wiley:00243590:media:lno10275:lno10275-math-0121 and urn:x-wiley:00243590:media:lno10275:lno10275-math-0122, Table 6). This is in accordance with differences in reactivity of iron oxide minerals of varying crystallinity to ΣH2S as reported in the literature (Canfield et al. 1992; Poulton et al. 2004).

Table 6. Reaction parameters used in the model
Parameter Value Units Source Range given by source (values used by source)
urn:x-wiley:00243590:media:lno10275:lno10275-math-0123 1.62 y−1 a 1.62
urn:x-wiley:00243590:media:lno10275:lno10275-math-0124 0.0086 y−1 a 0.0086
urn:x-wiley:00243590:media:lno10275:lno10275-math-0125 20 µM b 1–30, (20)
urn:x-wiley:00243590:media:lno10275:lno10275-math-0126 4 µM b 4–80, (2)
urn:x-wiley:00243590:media:lno10275:lno10275-math-0127 65 µmol g−1 b 65–100, (65, 100)
urn:x-wiley:00243590:media:lno10275:lno10275-math-0128 1.6 mM b 1.6, (1.6)
urn:x-wiley:00243590:media:lno10275:lno10275-math-0129 5000 mM−1 y−1 b 10,000, (5000)
urn:x-wiley:00243590:media:lno10275:lno10275-math-0130 140,000 mM−1 y−1 b 140,000, (140,000)
urn:x-wiley:00243590:media:lno10275:lno10275-math-0131 300 mM−1 y−1 b 300, (300)
urn:x-wiley:00243590:media:lno10275:lno10275-math-0132 1 mM−1 y−1 b 1
urn:x-wiley:00243590:media:lno10275:lno10275-math-0133 160 mM−1 y−1 b ≥160, (160)
urn:x-wiley:00243590:media:lno10275:lno10275-math-0134 107 mM−1 y−1 b 107, (107)
urn:x-wiley:00243590:media:lno10275:lno10275-math-0135 8 mM−1 y−1 b ≤100, (8)
urn:x-wiley:00243590:media:lno10275:lno10275-math-0136 5.6 mM−1 y−1 b ≤100, (8)
k8 1000 mM−1 y−1 Model constrained
k9 120 mM−1 y−1 b 10, (10)
k10 0.0074 mM−1 y−1 Model constrained
k11 3 y−1 c 3, (3)
k12 0.001 mM−1 y−1 Model constrained
k13 0.6 y−1 c 0.6, (0.6)
k14 0.51 mM−1 y−1 Model constrained
k15 0. 39 mM−1 y−1 Model constrained
k16 0.001 mM−1 y−1 Model constrained
k17 0.047 mM−1 y−1 Model constrained
  • Sources: (a) Reed et al. (2011); (b) Wang and Van Cappellen (1996); (c) Berg et al. (2003).

The model was run to steady state for 110 y and temporal changes in OM and Fe(OH)3 loading and sediment accumulation rates were then adjusted to fit the data (Fig. 1). The resulting scenario for this site is as follows: Sedimentation rates were initially high at 0.50 g cm−2 y−1 around 1970, reached a maximum of 0.63 g cm−2 y−1 between the years 1992 and 2002 and then rapidly decreased during the last 10 y of the simulation (0.03 g cm−2 y−1). Before the onset of anthropogenic eutrophication, the influx of organic carbon was 6.2 mol m−2 y−1 (Fig. 1), which falls within the range of 1.1 to 8.2 mol m−2 y−1 estimated previously for Bothnian Sea sediments (Algesten et al. 2006). The OM loading then began to increase, followed by a large OM pulse during the late 1990s and early 2000s and lower inputs during the last 10 y (Fig. 1). Fe(OH)3 fluxes increased during the late 1990s and early 2000s and subsequently decreased during the last 10 y (Fig. 1), consistent with changes in OM loading.

Details are in the caption following the image

Time-dependent sedimentation rate and fluxes of OM and Fe(OH)3 at the sediment-water interface as implemented in the baseline scenario.

The transient evolution of OM loading reported here compares well to qualitative trends reported in previous studies on eutrophication in the Bothnian Sea (Fleming-Lehtinen et al. 2008; Fleming-Lehtinen and Laamanen 2012). No quantitative data are available to further constrain our scenario. However, we note that, based on the location of our study site in a deep basin, it is likely that major shifts in lateral input of OM and FeOH3 and sediment focusing play a key role in controlling the strong temporal changes. Eutrophication-induced temporal changes in lateral transfer of iron from shallow areas to deep basins have been well-described for other Baltic Sea basins (Fehr et al. 2010; Jilbert and Slomp 2013; Lenz et al. 2015).

The transient scenario described above (“baseline scenario”) is subsequently used in a sensitivity analysis. In this analysis, the effects of variations in single parameters on the system were evaluated by running the scenario again but in this case setting the OM and Fe(OH)3 loadings, the O2 and urn:x-wiley:00243590:media:lno10275:lno10275-math-0137 bottom-water concentrations, and the rate parameter of Fe-AOM (k10, Table 6) to 80 and 120% of the original values (specified in the baseline scenario). To assess the role of environmental forcings in more detail, the OM loading was also changed simultaneously with the Fe(OH)3 loading and urn:x-wiley:00243590:media:lno10275:lno10275-math-0138 bottom-water concentration.

Results

The modeled and measured depth profiles of the various porewater constituents (CH4, urn:x-wiley:00243590:media:lno10275:lno10275-math-0139, urn:x-wiley:00243590:media:lno10275:lno10275-math-0140, ΣH2S, and Fe2+) for the year 2012 are similar and in line with an SMTZ that is located very close to the SWI (Fig. 2). Porewater urn:x-wiley:00243590:media:lno10275:lno10275-math-0141 and ΣH2S are only present in the upper 12 cm of sediment. A strong enrichment of total S and depletion of Fe(OH)3 is observed in the SMTZ. Below 12 cm depth, Fe2+ builds up in the absence of ∑H2S and accumulates to 1.8 mM. The reductive dissolution of Fe(OH)3 by ΣH2S results in a release of urn:x-wiley:00243590:media:lno10275:lno10275-math-0142 into the porewater around the SMTZ. Part of the liberated urn:x-wiley:00243590:media:lno10275:lno10275-math-0143 diffuses downwards below the SMTZ where it precipitates with Fe2+ to form authigenic vivianite, a reduced iron phosphate mineral. Some of the urn:x-wiley:00243590:media:lno10275:lno10275-math-0144 becomes readsorbed to Fe(OH)3 (Supporting Information Fig. S1). Porewater Fe2+ also precipitates as siderite (Fig. 2).

Details are in the caption following the image

Depth profiles of key porewater and solid phase constituents for a location in the Bothnian Sea for 2012 (site US5B, water depth 214 m; Egger et al. 2015a, 2015b) plotted along with best model fits.

The temporal trends in the porewater and solid phase constituents as modeled for 1970 to 2012 reveal a highly dynamic system (Fig. 3). The increase in OM loading between 1970 and 2002 has a strong impact on the urn:x-wiley:00243590:media:lno10275:lno10275-math-0145 penetration depth, which decreases from > 50 cm in 1970 to < 5 cm in 2002. urn:x-wiley:00243590:media:lno10275:lno10275-math-0146 and CH4 concentrations accumulate in the porewater. From the year 2000 onwards, ∑H2S emerges in the porewater and significantly more sulfide is stored in solid phases. Concentrations of Fe(OH)3 at depth decrease significantly between 1990 and 2012, while dissolved Fe2+ and siderite accumulate.

Details are in the caption following the image

Time-dependent evolution of key porewater and solid phase constituents with depth as calculated for the baseline scenario (1970–2012).

Following the onset of more eutrophic conditions, the organoclastic Fe(OH)3 and urn:x-wiley:00243590:media:lno10275:lno10275-math-0147 reduction rates shoal and intensify (Fig. 4). Methanogenesis occurs thoughout the sediment domain and its importance increases with time. The oxidation of CH4 with O2, Fe(OH)3, and urn:x-wiley:00243590:media:lno10275:lno10275-math-0148 intensifies and the reaction fronts move upwards toward the SWI. After 2002, the CH4 oxidation rates decrease and reaction fronts move to slightly greater depth.

Details are in the caption following the image

Time-dependent reaction profiles with depth as calculated for the baseline scenario (1970–2012). Rates are in mol C m−3 y−1.

Depth-integrated rates of transformations of CH4 and iron in the sediment for 2012 suggest that nearly all CH4 produced in the sediment is oxidized (Fig. 5). Fe-AOM accounts for 9% of total CH4 oxidation, while urn:x-wiley:00243590:media:lno10275:lno10275-math-0149 and O2 account for 90% and 1%, respectively. Iron cycling is strongly affected by Fe-AOM (Fig. 5), as it accounts for 46% of Fe(OH)3 reduction (Fig. 5). A significant proportion of the porewater Fe2+ precipitates as either FeSx, siderite, or vivianite and is subsequently buried.

Details are in the caption following the image

Budgets of carbon and iron for the year 2012 in the baseline scenario. The fluxes in parentheses are in mol C m−2 y−1 and mol Fe m−2 y−1 in the left and right panels, respectively.

Changing the rate of reactive OM deposition within 20% of the model calibration has a noticeable effect on the CH4 (Fig. 6a). A lower OM flux leads to deeper urn:x-wiley:00243590:media:lno10275:lno10275-math-0150 penetration and less ΣH2S and solid sulfur (S) accumulation in the sediment, while the Fe(OH)3 concentrations increase. Conversely, a higher OM loading leads to a steeper urn:x-wiley:00243590:media:lno10275:lno10275-math-0151 gradient, a significantly larger ΣH2S peak, and the depletion of Fe(OH)3 at the same depth. Changing the Fe(OH)3 loading has a moderate effect on the CH4 profile (Fig. 6b). The ΣH2S peak and Fe(OH)3 concentration at that depth are very sensitive, however, with a lower Fe(OH)3 loading leading to a much larger ΣH2S peak and lower Fe(OH)3, while increasing the Fe(OH)3 loading has the opposite effect. Varying the bottom-water O2 concentration has a negligible effect on the porewater profiles (Fig. 6c). Increasing the bottom-water concentration of urn:x-wiley:00243590:media:lno10275:lno10275-math-0152 results in slightly less accumulation of CH4, a smaller and deeper ΣH2S peak, and a different shape of the Fe(OH)3 profile. Table 7 lists the integrated reaction rates corresponding to the sensitivity analyses shown in Fig. 6.

Details are in the caption following the image

Model sensitivity analyses illustrated in profiles of organic carbon, CH4, urn:x-wiley:00243590:media:lno10275:lno10275-math-0153, ΣH2S, Fe2+, Fe(OH)3, and S solids for the year 2012. The model sensitivity to changing upper boundary conditions: (a) OM loading, (b) Fe(OH)3 loading, (c) O2 bottom-water concentration, and (d) urn:x-wiley:00243590:media:lno10275:lno10275-math-0154 bottom-water concentration (solid line = baseline, dotted line = 20% increase, dashed line = 20% decrease).

Table 7. Sensitivity analysis: Depth integrated reaction rates (mol C m−2 y−1)
O2 + CH4 urn:x-wiley:00243590:media:lno10275:lno10275-math-0155 + OM SO4-AOM urn:x-wiley:00243590:media:lno10275:lno10275-math-0156 Fe-AOM
Baseline 0.011 0.137 0.787 0.003 0.080
urn:x-wiley:00243590:media:lno10275:lno10275-math-0157 0.004 0.144 0.607 0.003 0.065
urn:x-wiley:00243590:media:lno10275:lno10275-math-0158 0.020 0.150 0.981 0.005 0.085
urn:x-wiley:00243590:media:lno10275:lno10275-math-0159 0.010 0.146 0.820 0.004 0.058
urn:x-wiley:00243590:media:lno10275:lno10275-math-0160 0.009 0.141 0.765 0.004 0.099
urn:x-wiley:00243590:media:lno10275:lno10275-math-0161 0.010 0.150 0.793 0.004 0.079
urn:x-wiley:00243590:media:lno10275:lno10275-math-0162 0.011 0.127 0.781 0.003 0.081
urn:x-wiley:00243590:media:lno10275:lno10275-math-0163 0.011 0.136 0.800 0.004 0.070
urn:x-wiley:00243590:media:lno10275:lno10275-math-0164 0.010 0.139 0.776 0.003 0.087
urn:x-wiley:00243590:media:lno10275:lno10275-math-0165 0.030 0.098 0.788 0.004 0.087
urn:x-wiley:00243590:media:lno10275:lno10275-math-0166 0.004 0.188 0.773 0.003 0.072
urn:x-wiley:00243590:media:lno10275:lno10275-math-0167 0.001 1.055 0.406 0.003 0.001

Combining the response of the geochemical system in terms of reaction rates, chemical profiles and sediment effluxes allows a mechanistic, process-based understanding of the controls on Fe-AOM and sediment-water exchange of Fe2+ under varying environmental conditions. Given that the system is most sensitive to changes in Fe(OH)3 and OM loading (Fig. 6), only the results of the sensitivity analyses where these boundary conditions are varied are discussed in further detail. Increasing the OM loading leads to slightly higher Fe-AOM rates and methanogenesis rates. Almost exclusively the β phase of Fe(OH)3 is involved in Fe-AOM (Fig. 7). A higher Fe(OH)3 loading increases the amount of Fe2+ produced during Fe-AOM, but has a relatively small effect on carbon cycling in the sediment (Fig. 7).

Details are in the caption following the image

Integrated areal reaction rates and fluxes of Fe2+ (top) and CH4 (bottom) under different OM (left) and FeOH3 (right) fluxes to the sediment-water interface (black = baseline, white = 20% increase, gray = 20% decrease) for the year 2012.

The response of the depth integrated rate of Fe-AOM for the year 2012 to a wider range of the key forcings (Fig. 8a,c) shows that the overall reaction rate increases with higher Fe(OH)3 loadings and lower urn:x-wiley:00243590:media:lno10275:lno10275-math-0168 concentrations. The effect of OM adds a nonlinearity to the system response, as there appears to be an optimal amount of OM input for maximum Fe-AOM intensity, at a given Fe(OH)3 flux or urn:x-wiley:00243590:media:lno10275:lno10275-math-0169 concentration, with both higher and lower fluxes of OM resulting in lower overall rates of Fe-AOM. Increasing the bottom-water urn:x-wiley:00243590:media:lno10275:lno10275-math-0170 concentration strongly impedes Fe-AOM rates (Fig. 8c), and urn:x-wiley:00243590:media:lno10275:lno10275-math-0171 concentrations above 14 mM eliminate the Fe-AOM reaction (Table 7).

Details are in the caption following the image

Depth integrated rates of Fe-AOM and Fe2+ efflux out of the sediment for varying OM and Fe(OH)3 fluxes to the sediment-water interface and urn:x-wiley:00243590:media:lno10275:lno10275-math-0172 bottom-water concentrations for the year 2012. The fluxes and urn:x-wiley:00243590:media:lno10275:lno10275-math-0173 concentrations are expressed as a percentage relative to the value in the baseline scenario, i.e., the baseline is zero, a twofold higher flux or concentration is 100% and a zero flux or concentration is −100%.

The modeled Fe2+ effluxes for 2012 are negligible. They increase, however, with increasing OM loading and also show a nonlinearity in the response to the Fe(OH)3 flux (Fig. 8b). Lower urn:x-wiley:00243590:media:lno10275:lno10275-math-0174 bottom-water concentrations, representing almost fresh water, can lead to order of magnitude higher Fe2+ effluxes (Fig. 8d).

Discussion

Coastal eutrophication and Fe-AOM

The model results presented in this study confirm that increased inputs of OM to the sediment can induce an upward vertical migration of the SMTZ in coastal sediments, as suggested previously for this site (Slomp et al. 2013; Egger et al. 2015a, 2015b). Our model results show that increased rates of urn:x-wiley:00243590:media:lno10275:lno10275-math-0175 reduction and methanogenesis linked to enhanced inputs of OM (Fig. 1) can initiate a relatively rapid upward movement of the SMTZ (Fig. 3) thus bringing it close to the sediment surface. Reduced inputs of OM over the last 10 y have now brought the upward displacement of the SMTZ to a halt. The fixation of the SMTZ at its current position has resulted in the accumulation of authigenic FeSx minerals in a relatively narrow zone over the past decade (Fig. 3).

High sediment accumulation rates and Fe(OH)3 loading have allowed a significant portion of the Fe(OH)3 to become buried below the newly established SMTZ. The preservation of reducible Fe(OH)3 below the zone of urn:x-wiley:00243590:media:lno10275:lno10275-math-0176 reduction has stimulated Fe-AOM, which has led to the build up of dissolved Fe2+ in the porewater below the SMTZ. Our results support earlier suggestions by Egger et al. (2015a) that anthropogenic eutrophication of coastal systems may increase the importance of Fe-AOM in coastal environments by inducing nonsteady state diagenesis.

Controls on the Fe-AOM reaction

Besides the quest for conclusive evidence that AOM is truly coupled to Fe(OH)3 reduction in natural environments and for the microorganisms carrying out the reaction, there is also the question what processes could potentially control in-situ rates of Fe-AOM in sediments. While laboratory studies suggest that the availability Fe(OH)3 can boost the rate of AOM (Beal et al. 2009; Sivan et al. 2014; Egger et al. 2015a) and that the addition of multiple electron acceptors may decrease the AOM rate (Segarra et al. 2013), the potential response of Fe-AOM to variations in environmental conditions in the field has so far not yet been described.

Here, we calibrated a transient sedimentary model to data from a coastal brackish basin where various lines of evidence suggest that Fe-AOM occurs (Egger et al. 2015a). The model assumes a direct coupling between Fe(OH)3 reduction and CH4 oxidation as proposed by Beal et al. (2009) (see Eq. 1), rather than indirect iron stimulated SO4-AOM (Sivan et al. 2014) because ΣH2S and urn:x-wiley:00243590:media:lno10275:lno10275-math-0177 are undetectable below the SMTZ. Starting from this assumption, variations in key environmental forcings can be implemented in the model to discern potential controls on Fe-AOM.

The model assumes three electron acceptors which can react with the available CH4: O2, urn:x-wiley:00243590:media:lno10275:lno10275-math-0178, and Fe(OH)3. Our results show that Fe-AOM is mostly restricted to parts of the sediment where other electron acceptors than Fe(OH)3 are absent. The oxidation of CH4 coupled to O2 is by far the most favorable kinetically (k5 in comparison to k9 and k10, Table 6), but is constrained to the upper millimeters of sediment where O2 is present. The rate constant for SO4-AOM is five orders of magnitude lower than that of aerobic CH4 oxidation, but occurs over a larger depth interval because urn:x-wiley:00243590:media:lno10275:lno10275-math-0179 is present throughout the upper 10 cm. In the model, the kinetics of Fe-AOM follow the same bimolecular rate law as the other CH4 oxidation pathways. CH4 in the top is oxidized by O2 and urn:x-wiley:00243590:media:lno10275:lno10275-math-0180, while the Fe-AOM reaction is only optimal at depths where urn:x-wiley:00243590:media:lno10275:lno10275-math-0181 is depleted (compare Figs. 3, 4).

Enhanced bottom-water urn:x-wiley:00243590:media:lno10275:lno10275-math-0182 concentrations significantly reduce Fe-AOM rates as less CH4 is available for Fe-AOM. This is due to enhanced organoclastic urn:x-wiley:00243590:media:lno10275:lno10275-math-0183 reduction, which suppresses methanogenesis. For example, a three-fold increase of the urn:x-wiley:00243590:media:lno10275:lno10275-math-0184 concentration (corresponding to 14.5 mM urn:x-wiley:00243590:media:lno10275:lno10275-math-0185) leads to a CH4 oxidation rate that is only 46% of that in the baseline scenario and a Fe-AOM rate near zero (Table 7).

The availability of Fe(OH)3 for AOM depends on the fraction of the sedimentary Fe(OH)3 that can react with OM and ΣH2S. A higher OM loading leads to a steeper urn:x-wiley:00243590:media:lno10275:lno10275-math-0186 gradient (Fig. 6a) and thus more diffusion of urn:x-wiley:00243590:media:lno10275:lno10275-math-0187 into the sediment. This enhances ΣH2S production and subsequent Fe(OH)3 scavenging by ΣH2S. Therefore, systems that receive high reactive OM fluxes but no associated high Fe(OH)3 flux, will not be suitable environments for Fe-AOM in the surface sediment. The accumulation of Fe2+ can only occur when the Fe(OH)3 loading is high enough to constrain the ΣH2S to the upper part of the sediment. Figure 8a shows that almost no Fe-AOM was observed in simulations with a 50% lower Fe(OH)3 loading.

The impact of changes in Fe(OH)3 loading on OM degradation is generally limited as the organoclastic urn:x-wiley:00243590:media:lno10275:lno10275-math-0188 reduction rate is much larger (Table 7). As the methanogenesis rate is dependent on the total sum of the other OM degradation pathways (Table 4), organoclastic Fe(OH)3 reduction has a fairly small impact on the in-situ CH4 production. In line with Egger et al. (2015a), we find that Fe-AOM only accounts for a small fraction of the total CH4 oxidation rate, and therefore has a relatively small effect on the CH4 profile.

Signatures of Fe-AOM

As previously outlined, direct identification of the Fe-AOM process in natural environments is difficult because of the lack of unequivocal geochemical signatures. It is therefore useful to test how the geochemical profiles of the coupled iron-sulfur-carbon cycles can serve as signatures for the presence of Fe-AOM. By simulating a range of environments, we have constructed a group of theoretical sedimentary settings that we can use to assess how the presence of Fe-AOM is reflected in the geochemical profiles of its associated species.

The usefulness of CH4 profiles for identifying Fe-AOM is limited, as the simulation outcomes show that CH4 is not very sensitive to different intensities of Fe-AOM. In the simulation with a 20% higher Fe(OH)3 loading CH4 concentrations were slightly lower than obtained in calculations with the baseline scenario. This was not only caused by the increased Fe-AOM rates, but also by the higher rates of organoclastic Fe(OH)3 reduction (Table 7). The latter aspect affects the partitioning of OM degradation among electron acceptors and slightly suppresses methanogenesis. Fe-AOM leads to the production of large quantities of Fe2+, which can lead to the accumulation of Fe2+ at depth. However, to interpret the impact of Fe-AOM on Fe2+ profiles, the sinks of Fe2+ need to be taken into account (Fig. 7). Most dissolved Fe2+ is removed through FeSx formation at a rate that is dependent on multiple factors. The availability of ΣH2S is controlled by the urn:x-wiley:00243590:media:lno10275:lno10275-math-0189 reduction rate, which in turn is dependent on the urn:x-wiley:00243590:media:lno10275:lno10275-math-0190 bottom-water concentration and the OM loading. The depth and breadth of the Fe-AOM front is consequential, as Fe2+ produced higher up in the sediment is more likely to react with ΣH2S. Also a significant part of the Fe2+ is stored in carbonate and phosphate minerals, i.e., siderite and vivianite (Fig. 7). In the simulations for the baseline scenario, Fe2+ does not accumulate at depth before 1982 (Fig. 3), while the total Fe-AOM rates over depth are significant (Fig. 4). Therefore, Fe2+ concentration profiles are not a robust signature of Fe-AOM.

Conversely, Fe2+ accumulation at depth, may signal the occurrence of Fe-AOM, as it can explain the concurrent production of CH4 and Fe2+. Such a simultaneous production is hard to explain otherwise if the assumption is correct that methanogens are generally outcompeted for common substrates by organoclastic Fe-reducers (Egger et al. 2015a and references therein). In the simulations, Fe-AOM is the only process that can produce Fe2+ at depth. The other Fe2+ producing reactions are dependent on either O2, ∑H2S, or the alpha phase of Fe(OH)3, which are only available in the upper few centimeters of sediment (see Figs. 3, 4). To achieve deep Fe2+ production through Fe-AOM in our scenario, the Fe(OH)3 needs to penetrate to sufficient depth. This requires a fraction of the Fe(OH)3 pool to be unreactive toward OM, which is in line with the characteristics of the Fe urn:x-wiley:00243590:media:lno10275:lno10275-math-0191 phase. Mass balance calculations show that the Fe urn:x-wiley:00243590:media:lno10275:lno10275-math-0192 phase is almost exclusively involved in the Fe-AOM reaction (Fig. 7).

Environmental settings

Our results suggest that Fe-AOM is promoted in brackish coastal marine sediments with relatively high OM fluxes and Fe(OH)3 loadings. Estuarine environments such as Chesapeake Bay, Long Island Sound and the Scheldt estuary all exhibit these conditions. Thus they are all potential candidate sites for investigating the occurrence and scale of AOM coupled to metal oxide reduction. From our sensitivity analyses we see that higher bottom-water urn:x-wiley:00243590:media:lno10275:lno10275-math-0193 concentrations significantly reduce the probability of the occurrence and the intensity of Fe-AOM (Fig. 8). This indicates that the occurrence of this process in surface sediments is, in principle, restricted to low salinity environments. However, Fe-AOM can also occur in deep-sea sediments with typical bottom-water urn:x-wiley:00243590:media:lno10275:lno10275-math-0194 concentrations for marine settings (∼28 mM) if there is a low sulfate zone at depth, as shown by Riedinger et al. (2014). At their site, CH4 diffuses up from greater depth and Fe-AOM is believed to occur at 6–8 m depth below the SWI. In our model simulations, higher urn:x-wiley:00243590:media:lno10275:lno10275-math-0195 bottom-water concentrations reduce Fe-AOM rates primarily by suppressing methanogenesis. However, higher bottom-water urn:x-wiley:00243590:media:lno10275:lno10275-math-0196 concentrations will be less consequential for Fe-AOM when Fe(OH)3 is buried to greater depth (outside our model domain) under oligotrophic conditions. At depth, the CH4 can be formed in situ when the system shifts from oligotrophic to eutrophic conditions, which leads to an upward movement of the SMTZ and methanogenic zone. There can also be an ex situ deep CH4 source, which explains why, for example, cold seep environments are also key settings for Fe-AOM (Beal et al. 2009; Wankel et al. 2012).

A necessary precondition for Fe-AOM, is the presence of sufficient amounts of Fe(OH)3 to react with the available CH4, as there is a 8 : 1 stoichiometric ratio of Fe(OH)3 to CH4 in the net reaction. We show that relatively high fluxes of Fe(OH)3 are necessary to provide enough material for the reaction to occur. Furthermore, the simulations indicate that Fe-AOM is favored in the presence of less reactive iron phases that can penetrate deeper into the sediment. Settings, which receive high amounts of Fe(OH)3, such as coastal systems near large rivers (Poulton and Raiswell 2002), are therefore also candidates for hosting the reaction.

The impact of OM loading on Fe-AOM is complicated as it affects the depth penetration of urn:x-wiley:00243590:media:lno10275:lno10275-math-0197, the urn:x-wiley:00243590:media:lno10275:lno10275-math-0198 reduction rate, the rate of methanogenesis, and the persistence of Fe(OH)3 at depth. The sensitivity analyses show that a wide range of OM loadings may lead to substantial Fe-AOM rates (Fig. 8), but that too much and too little OM can have a detrimental effect on the process.

In summary, the optimal environments for the Fe-AOM process are sedimentary settings which are depositional and receive sufficient OM loading (much of which can be refractory), receiving high influxes of Fe(OH)3 (preferably of a more refractory nature), and which have sufficiently low salinity. Coastal areas, with brackish water columns and a connection to rivers are thus ideal environments for the Fe-AOM process. In more marine settings, Fe-AOM is far less likely, but may still occur, provided there is an ex-situ source of CH4 and enough Fe(OH)3 to fuel the reaction.

Sediment-water exchange fluxes of Fe2+

Continued shuttling of iron through repeated reduction-oxidation cycles and lateral iron transport on oxic and hypoxic continental shelves depends on the replenishment of iron from near-coastal regions. Our results suggest that, besides ∑H2S-driven and organoclastic Fe(OH)3 reduction, Fe-AOM could contribute to benthic release of Fe2+ and thus to the supply of iron from estuarine sediments to continental shelves. In the model result for 2012, the Fe2+ efflux to the overlying water is negligible. However, the results for the sensitivity analyses show that increased OM loadings can enhance exchange fluxes of Fe2+ to the overlying water (Fig. 8), with values of up to 0.04 mol m−2 y−1 being reached. These Fe2+ fluxes fall within the low end of the typical range estimated for continental margin sediments overlain by oxic waters (0 to ca. 0.12 mol m−2 y−1; e.g., Severmann et al. 2010; Dale et al. 2015). The increased release of Fe2+ on increased OM loading is the result of a shift in the zones of Fe-AOM and organoclastic Fe(OH)3 reduction toward the SWI (Fig. 4). Increased bottom-water urn:x-wiley:00243590:media:lno10275:lno10275-math-0199 concentrations hinder Fe2+ effluxes due to lower rates of Fe(OH)3 reduction (Fig. 8) and more sulfidic scavenging of Fe2+. Very low bottom-water urn:x-wiley:00243590:media:lno10275:lno10275-math-0200 concentrations and high OM loadings may lead to Fe2+ effluxes > 0.12 mol m−2 y−1, which suggests that Fe-AOM has the potential to greatly impact iron dynamics in low-salinity coastal environments.

Acknowledgments

IT, ME, and CPS acknowledge funding by the Netherlands Organisation for Scientific Research (NWO Vici grant) and the European Research Council (ERC; Starting Grant #278364). This work was carried out under the program of the Netherlands Earth System Science Centre (NESSC), financially supported by the Ministry of Education, Culture and Science (OCW).