Vous êtes sur la page 1sur 20

Science of the Total Environment 649 (2019) 264–283

Contents lists available at ScienceDirect

Science of the Total Environment

journal homepage: www.elsevier.com/locate/scitotenv

Iron sulfide formation in young and rapidly-deposited permeable sands


at the land-sea transition zone
Stephan L. Seibert a,⁎, Michael E. Böttcher b, Florian Schubert b, Thomas Pollmann c, Luise Giani c,
Sumiko Tsukamoto d, Manfred Frechen d, Holger Freund e, Hannelore Waska f, Heike Simon f, Tobias Holt a,
Janek Greskowiak a, Gudrun Massmann a
a
Hydrogeology and Landscape Hydrology Group, Institute for Biology and Environmental Sciences, Carl von Ossietzky University of Oldenburg, Ammerländer Heerstraße 114-118, D-26129 Olden-
burg, Germany
b
Geochemistry and Isotope Biogeochemistry Group, Department of Marine Geology, Leibniz Institute for Baltic Sea Research (IOW), D-18119 Warnemünde, Germany
c
Soil Science Group, Institute for Biology and Environmental Sciences, Carl von Ossietzky University of Oldenburg, Ammerländer Heerstraße 114-118, D-26129 Oldenburg, Germany
d
Leibniz Institute for Applied Geophysics (LIAG), S3: Geochronology and Isotope Hydrology, Stilleweg 2, D-30655 Hannover, Germany
e
Geoecology Group, Institute for Chemistry and Biology of the Marine Environment (ICBM), Carl von Ossietzky University of Oldenburg, Schleusenstr. 1, D-26382 Wilhelmshaven, Germany
f
Research Group for Marine Geochemistry (ICBM-MPI Bridging Group), Institute for Chemistry and Biology of the Marine Environment (ICBM), Carl von Ossietzky University of Oldenburg, D-
26129 Oldenburg, Germany

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Biogeochemical cycles in coastal


organic-poor, permeable sands are not
well studied.
• Our multidisciplinary approach covers
aspects of coastal litho- and hydro-
spheres.
• Methods include hydrogeochemical,
sulfur isotope, DOM, and OSL age inves-
tigations.
• SRR are extremely low and organic car-
bon limits iron sulfide formation.
• Biogeochemical cycles are linked to
(seasonal) hydrochemical variations.

a r t i c l e i n f o a b s t r a c t

Article history: Organic-poor, permeable quartz sands are often present at land-sea transition zones in coastal regions. Yet, the
Received 6 June 2018 biogeochemical cycles of carbon, sulfur, and iron are not well studied here. The aim of this work was, therefore,
Received in revised form 20 August 2018 to improve our understanding regarding the chemical processes in these prominent coastal sediments. A 10 m
Accepted 20 August 2018
core was collected at a dune base of the barrier island Spiekeroog, Germany, for this purpose. Additionally,
Available online 23 August 2018
groundwater was sampled from a multi-level well for one year to record seasonal hydrochemical variations.
Editor: José Virgílio Cruz Methods included the analyses of geochemical (total carbon, total inorganic carbon, reactive iron, total sulfur, re-
duced inorganic sulfur) and hydrochemical parameters (field parameters, major ions, DOC, and molecular com-
Keywords: positions of DOM), as well as stable sulfur isotopes (δ34S-sulfate, -sulfide, -total reduced inorganic sulfur).
Pyrite formation Moreover, optically stimulated luminescence (OSL) dating was applied. Results show that the core sediments
Stable sulfur isotopes are very young (b500 a) and were rapidly deposited. They are characterized by remarkably low contents of or-
OSL dating ganic carbon (b0.1% dw.), reactive iron (~10 mmol/kg), and iron sulfides (b3 mmol/kg). Groundwater salinities
Barrier island were low in the top core sediments and increased at depth during most times of the year. However, the sampling
Spiekeroog Island
site is subject to (seasonally) varying salinities, which could be linked to the biogeochemical cycles. For instance,
Subterranean estuary

⁎ Corresponding author.
E-mail address: stephan.seibert@uni-oldenburg.de (S.L. Seibert).

https://doi.org/10.1016/j.scitotenv.2018.08.278
0048-9697/© 2018 Elsevier B.V. All rights reserved.
S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283 265

the infiltration of seawater-derived labile DOM during inundation events drives microbial respiration besides
sedimentary organic matter. Oxygen and nitrate were the dominant electron acceptors for the decomposition
of organic matter in near-surface groundwater, while sulfate reduction was constrained to the lower brackish
sediments. Here, authigenic pyrite formation was inferred based on the detection of dissolved sulfide, intact py-
rite framboids, and matching stable sulfur isotope signatures of dissolved and solid sulfides. We concluded that
the extremely low organic carbon contents limit pyrite formation in the organic-poor, permeable quartz sands.
© 2018 Elsevier B.V. All rights reserved.

1. Introduction the marine environment were previously attributed to a cycling of sul-


fur species, including re-oxidation of dissolved sulfide and subsequent
The environmental cycles of carbon, sulfur and iron are tightly linked disproportionation of sulfur intermediates (Canfield and Thamdrup,
to each other. Organic matter is constantly produced, degraded, and de- 1994; Habicht et al., 1998; Habicht and Canfield, 2001). More recent
posited in all marine and terrestrial ecosystems (e.g., Burdige, 2007). studies, however, demonstrated that high isotope discrimination may
Sulfur in the form of sulfate is a major constituent in seawater also be achieved by dissimilatory sulfate reduction alone (Wortmann
(Howarth, 1984) and present in the atmosphere. Furthermore, iron is et al., 2001; Rudnicki et al., 2001; Canfield et al., 2010; Sim et al.,
delivered to the sea as nanoparticulate (oxyhdr)oxides (e.g., Moore 2011) and that the reactivity of the electron donors may be involved
and Braucher, 2008; Raiswell et al., 2008; Tagliabue et al., 2009; in the developing sulfur isotope discrimination (Kaplan and
Raiswell and Canfield, 2012). Rittenberg, 1964; Sim et al., 2011).
In (marine) sediments, redox conditions rapidly become reducing in Mineral structures of iron sulfides present a further means to inves-
the presence of reactive organic matter (e.g., Canfield et al., 1993a; tigate chemical conditions during formation (e.g., Wilkin et al., 1996;
Thamdrup et al., 1994; Al-Raei et al., 2009), resulting in the anaerobic Böttcher and Lepland, 2000; Merinero et al., 2009; Lin et al., 2016). As
oxidation of organic carbon using mainly sulfate as electron acceptor an example, size distribution and variability of pyrite framboids were
(Jørgensen, 1982). The reduction of metal oxides (i.e., iron (oxyhdr)ox- successfully used to infer the redox conditions during sediment deposi-
ides and manganese oxides) is often less relevant for the bulk oxidation tion or to distinguish between syngenetic and diagenetic pyrite forma-
of organic carbon and may proceed via microbes (e.g., Nealson and tion (e.g., Wilkin et al., 1996).
Myers, 1990; Lovley, 1991) or coupled to the chemical oxidation of re- The intense study of isotope biogeochemistry in coastal environ-
duced sulfide species (e.g., Canfield, 1989; Thamdrup et al., 1994; ments has substantially improved our understanding of carbon, sulfur,
Gagnon et al., 1995; Blonder et al., 2017). Methanogenesis classically and iron cycles in the past three decades. Studies were carried out in
presents the terminal stage in the redox sequence (Berner, 1981). How- fully/brackish marine (e.g., Canfield et al., 1993a; Thamdrup et al.,
ever, zones of iron (oxyhdr)oxide and sulfate reduction and 1994; Gagnon et al., 1995; Neretin et al., 2004; Neumann et al., 2005),
methanogenesis are often not strictly separated in nature and may over- (semi)euxinic (e.g., Lyons and Berner, 1992; Middelburg, 1991), estua-
lap, depending on iron (oxyhdr)oxide stability, sulfate concentrations, rine (e.g., Morse et al., 2007; Morgan et al., 2012; Kraal et al., 2013),
and substrate levels (e.g., Postma and Jakobsen, 1996; Jakobsen and marsh (e.g., Luther and Church, 1988; Dellwig et al., 2002), and fresh-
Postma, 1999). The role of sulfate as the prominent electron acceptor water environments (e.g., Marnette et al., 1993; Huerta-Diaz et al.,
in marine sediments was recently corroborated by Hansel et al. 1998; Canfield et al., 2010). However, very little is known about the cy-
(2015), who suggested that sulfate reduction may precede the cling of redox sensitive elements, including the formation of iron sul-
reduction of all types of iron (oxyhdr)oxides, including ferrihydrite, fides, in permeable sands with low organic carbon and reactive iron
even at low sulfate concentrations, a finding challenging classical, contents, as only few studies were carried out in such environments
thermodynamic-based redox models (e.g., Berner, 1981; Lovley (Böttcher et al., 1998, 2004; Jakobsen and Postma, 1999; Al-Raei et al.,
and Phillips, 1987; Lovley and Goodwin, 1988; Chapelle and Lovley, 2009).
1992). The aim of this study was to assess the cycling of carbon, sulfur, and
As reaction products of anaerobic organic matter degradation, sul- iron in organic-poor, permeable Holocene sands under the influence of
fide and ferrous iron are formed that may react to form iron sulfides fresh to brackish groundwater salinities. Investigations were carried out
like mackinawite, greigite, and finally pyrite (e.g., Berner, 1970; at the currently developing Ostplate, which forms the eastern part of the
Luther, 1991; Wilkin and Barnes, 1996; Passier et al., 1997; Rickard barrier island Spiekeroog, southern North Sea, Germany. Main objec-
and Luther, 1997; Neretin et al., 2004; Schoonen, 2004; Rickard tives were to study the geochemical conditions and limiting factors of
and Morse, 2005; Raiswell and Canfield, 2012). The formation/transfor- pyrite formation in such an environment and to assess the effect of vary-
mation processes of iron sulfides are inherently coupled to the availabil- ing salinity and redox gradients on the biogeochemical cycles. Interpre-
ity of reactive organic matter, sulfate, and iron (oxyhydr)oxides tations are based on a combination of geochemical, hydrochemical,
(e.g., Berner, 1970; Berner and Raiswell, 1984; Lyons and Berner, stable sulfur isotope, and mineral structure investigations, as well as op-
1992; Morse et al., 2007), and the application of geochemical indicators tically stimulated luminescence (OSL) dating.
(e.g., C/S ratio, degree of pyritization (DOP), acid volatile sulfide (AVS)/
pyrite ratio) proved useful in investigating the environmental condi- 2. Methodology
tions during their formation more closely (Raiswell and Canfield, 2012).
Stable sulfur isotope signatures of dissolved sulfate and sulfide as 2.1. Study site
well as solid iron sulfides have been widely used as additional tools to
study sulfur sources and biogeochemical processes (e.g., Passier et al., Core and groundwater sampling were carried out at the Ostplate,
1997; Böttcher et al., 2004; Kunzmann et al., 2017). This is applicable which is the eastern part of the barrier island Spiekeroog, Germany
because bacterially mediated sulfate reduction as well as oxidation of (Fig. 1a). Spiekeroog Island is located in the southern North Sea in
sulfides and disproportionation of sulfur intermediates are associated front of the North German coastline. It has a west-to-east extent of
with distinct isotope fractionations (Nakai and Jensen, 1964; Canfield ~9.8 km and a total surface area of ~21.3 km2 (Streif, 1990). To the
and Thamdrup, 1994; Habicht et al., 1998; Cypionka et al., 1998; east, the tidal inlet Harle separates it from the neighboring island
Böttcher et al., 2005). High stable sulfur isotope discrimination between Wangerooge. The local tidal range of the North Sea is 2.72 m, with a
dissolved sulfate and reduced sulfide of up to about 50‰ observed in mean high water of 1.39 m above sea level (masl) and a mean low
266 S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283

Fig. 1. a) Location of Spiekeroog Island at the North German coast line. Dotted areas present dune areas N3 masl after Röper et al. (2013), and the thick black lines indicate the positions of
the freshwater lens in the west (Tronicke, 1997) and the near-surface freshwater extent at the Ostplate (Holt et al., 2017), respectively. The red line indicates the cross-section A-A', where
core sampling site and groundwater monitoring wells are located. b) Cross section A-A'. The core sampling site (ground surface elevation = 3.18 masl) at monitoring site DN is indicated,
including the location of the four monitoring wells DN-a, -b, -c, and -d. The elevation profile was obtained from a digital elevation model of the year 2014 (Lower Saxony Water
Management, Coastal Defense and Nature Conservation Agency, NLWKN, 2014). Dotted vertical lines indicate the level of highest high water (hhw, 4.14 masl, Federal Waterways and
Shipping Administration (WSV, 2016)), and blue and red areas conceptualize the assumed extent of the freshwater lens during sampling in April 2016. Question marks indicate that
the exact position of the freshwater saltwater interface is unknown. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this
article.)

water of 1.33 m below sea level (mbsl; annual means from 1995 to 2.2. Core drilling and well construction
2015, Federal Waterways and Shipping Administration (WSV, 2016)).
While the oldest part of the west of Spiekeroog Island dates from Employing dry drilling and using a Nordmeyer DSB-3.4 L well rig, an
~1650 and was formed due to the aggregation of aeolian and marine undisturbed 10 m sediment core was collected at the Ostplate on 15/03/
sands (Sindowski, 1970), the Ostplate developed more recently as a 2016 during the construction of groundwater monitoring wells at the
consequence of land reclamation activities at the mainland. The latter northern dune base facing the North Sea (referred to as monitoring
resulted in a change of currents and the progressive filling and eastward site ‘DN’, Fig. 1b). The core was stored cool and dry in non-transparent
migration of the former Harle inlet. The filling of the inlet at the Ostplate casings until further processing. Three groundwater monitoring wells
proceeded via six distinguishable cycles of sand deposition that oc- (DN-a, -b, and -c) are situated in direct vicinity of the core sampling
curred after 1650 (cycles 1–4) and before 1600 (cycles 5–6), according site (b2 m distance), and one monitoring well (DN-d) was installed in
to Sindowski (1968). A prominent dune ridge with secondary dunes the bore of the collected core. All wells have a screen length of 1 m
started to form at the Ostplate around 1940. Progressively increasing each and are located at 10.1-9.1 (DN-d), 8.1-7.1 (DN-c), 5.3-4.3 (DN-
in width and height, it allowed for the formation of freshwater lenses b) and 2.0-1.0 (DN-a) m below ground surface (mbgs), respectively
(Röper et al., 2013; Fig. 1a). Holt et al. (2017) further assessed that the (Fig. 1b).
near-surface extent of these freshwater lenses is highly sensitive to sea-
sonal inundation events.
The upper 40 to 60 m of the geologic units in the Spiekeroog area 2.3. Sediment sampling and analysis
mainly consist of fine to coarse-grained, permeable quartz sands of Plio-
cene, Pleistocene, and Holocene origin, deposited under changing fluvia- Sediment sampling was carried out at the Luminescence Laboratory
tile, marine, and aeolian regimes (compare Röper et al., 2013; Seibert of the Leibniz Institute for Applied Geophysics (LIAG) in July 2016 and
et al., 2018). Minor sediment fractions comprise shell debris, low amounts included the collection of 8 and 31 subsamples for sediment OSL dating
of silt and clay, and local embeddings of peat at greater depth. The Holo- and geochemical analysis, respectively. Subsamples allotted for OSL dat-
cene base is located at ~14 to 18 mbsl at the study site (LBEG, 2018). ing were collected from ~0.5 to 7.4 m depth in ~100 cm intervals under
S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283 267

subdued red light. Additional subsamples for gamma spectrometry contents were corrected for remaining sulfate from pore waters
measurements for the determination of the radionuclide concentrations (Appendix A.3).
were collected at each depth. OSL dating could not be carried out for Grain size distributions were determined for five samples collected
core sediments at depths N7.4 mbgs, as the high contents of shell debris at 0.20, 1.80, 4.56, 7.75, and 9.48 mbgs by wet sieving and drying sample
impeded the sampling procedure. aliquots at 105 °C for 24 h. The mud fraction (b63 μm) of each sediment
For the geochemical analyses, two replicate samples were collected sample was obtained during a second round of wet sieving with a sep-
at sampling intervals of ~30 cm. The first replicates were stored in arate sample aliquot, followed by freeze-drying. For the characterization
60 cm3 centrifuge tubes and subsequently used for the determination of the mineralogy, texture, and chemical composition of solid phases in
of water content, sedimentary total carbon (TC), total inorganic carbon the separated mud fraction, we applied scanning electron microscopy
(TIC), reactive iron (Fedith, FeHCl), and total sulfur (TS) contents, grain coupled with an EDX system (MERLIN VP compact, Carl Zeiss with
size distributions, and microscopic analyses of sediment minerals. The AztecEnergy, Oxford Instruments). The mineral powder was first
second replicate samples were filled and preserved in 60 cm3 centrifuge scanned over night to detect and count iron sulfides, before the jump
tubes, containing 10 mL of a 20% ZnAc solution, and later used for the mode was used to analyze the texture and major element composition
determination of total reduced solid inorganic sulfur (TRIS) contents of the single phases in more detail.
and stable sulfur isotope analyses. All samples were immediately stored
in a refrigerator box and frozen within 4 to 6 h after sampling. 2.4. Groundwater sampling and analysis
The sediment water content of all samples was determined gravi-
metrically by weighing a sediment aliquot before and after freeze- Groundwater was sampled by means of a submersible pump
drying. All samples were sieved afterwards, and only the fraction (Eijkelkamp, Tauchpumpe “Gigant”) at all four monitoring wells (DN-
b1 mm was processed for geochemical analysis. The fraction N1 mm a, -b, -c, -d, Fig. 1b) during six sampling campaigns in April, June, Sep-
consisted of shell fragments only and was used to determine shell con- tember, and November 2016, as well as in January and March 2017 to
tents. TC, TIC, and TS were analyzed using an Eltra CS 580 device, and study seasonal variations of the hydrochemistry at the core sampling
total organic carbon (TOC) contents were obtained by subtracting TIC site. Field parameters (i.e., electrical conductivity (EC), redox potential,
from TC. Replicate measurements of TC were carried out for 13% (n = pH, dissolved oxygen concentrations, and temperature) were deter-
4) of the total sample number, and results were in very good agreement mined in the field, using a Hach HQ 40d multi device with sampling
with each other (average deviation b5%, Fig. A.1.1). probes. Sampling included the collection of samples for major ion, dis-
The buffered dithionite (Fedith.) as well as the cold 0.5 M hydro- solved organic carbon (DOC), alkalinity, and dissolved sulfide analyses.
chloric acid (FeHCl) extraction methods were used (Raiswell et al., A detailed description of the sampling procedures can be viewed in
1994) to extract the sedimentary reactive iron fraction available for Seibert et al. (2018), where hydrochemical data obtained from the
early diagenetic reaction with dissolved sulfide (e.g., Canfield, fresh upper two wells was partly presented. High sample volumes
1989). The citrate-dithionite method quantitatively leaches iron (~5 L per well) for the investigation of the stable sulfur isotope signa-
from reactive iron (oxyhydr)oxides (i.e., ferrihydrite, lepidocrocite, tures of dissolved sulfate (δ34S-SO2− 34
4 ) and dissolved sulfide (δ S-H2S)
goethite, and hematite) and some iron carbonates (e.g., siderite). were collected once in September 2016. Moreover, samples for the
The cold HCl method only extracts iron from poorly-crystalline iron characterization of dissolved organic matter (DOM) were collected in
(oxyhydr)oxides (e.g., ferrihydrite and some lepidocrocite), iron car- September 2016. In addition, DOM samples were collected at the
bonates, and some iron silicate phases (Canfield, 1989; Raiswell Ostplate from two monitoring wells of the nearby monitoring site ‘DS’
et al., 1994). Iron(II) derived from FeS is extracted by both methods. (fresh to brackish salinities) located at the southern side of the dune
Iron was determined as Fe2+ after reduction with hydroxylamine ridge, three monitoring wells of monitoring site ‘B' located at the
hydrochloride via spectrophotometry, using ferrozine as the beach (brackish to saline salinities), and seawater, respectively, in Sep-
complexing agent following Stookey (1970). Replicate measure- tember 2016 to study DOM compositions as a function of salinity (Ap-
ments of Fedith. were carried out for all samples and showed a good pendix 4).
overall agreement (Fig. A.2.1). To quantitatively fix the dissolved sulfide species for the δ34S analy-
The AVS (essentially mackinawite, FeS) and chromium reducible sis, groundwater was directly filled in 5 L canisters containing 100 mL of
sulfur (CRS; essentially pyrite, FeS2, and elemental sulfur, S0) fractions a 20% ZnAC solution, and the canisters were shaken immediately after-
in the sediments were determined for selected samples by a two-step wards. The precipitated ZnS was filtered off using a water-jet vacuum
sequential extraction (Fossing and Jørgensen, 1989), using 5 g sediment pump and 0.45 μm cellulose acetate filters, rinsed with distilled water,
aliquots. The AVS fraction was first extracted by the reaction with 1 M and dried in culture dishes 24 to 48 h after sample collection. The resid-
HCl for 1 h under a continuous stream of N2 gas. Potential H2S released ual, sulfide-free groundwater samples were stored in a refrigerator and
was quantitatively precipitated as ZnS, and concentrations were deter- used to quantitatively precipitate the dissolved sulfate, which was car-
mined spectrophotometrically with a Specord 40 spectrophotometer ried out within 2 to 3 weeks after sampling by heating each sample to
following the method of Cline (1969). The CRS fraction was extracted ~60 °C, decreasing the pH to ~2 to 3 with 37% HCl, and adding sufficient
from the remainder of the sample in a second step, using a hot acidic amounts of a 5% BaCl2 solution. The precipitated BaSO4 was filtered,
Cr(II)chloride solution. As for the AVS procedure, emerging H2S was rinsed, and dried after 24 to 48 h, as described for the ZnS precipitates.
trapped in ZnAc during the CRS distillation, and concentrations were de- In July 2017, additional BaSO4 precipitations were carried out for brack-
termined by spectrophotometry using the precipitated ZnS. TRIS (total ish groundwater samples collected in April and November 2016 and
reducible inorganic sulfur) is the sum of the AVS and CRS fractions. January and March 2017. Those samples had previously been frozen
AVS was not detected in any sample, and TRIS contents could not be after sampling and were only defrosted immediately before the BaSO4
quantified in sediments at ~2 to 7 mbgs, as no ZnS precipitation was ob- precipitation procedure.
served. For the δ34S analysis of sedimentary TRIS, the CRS fraction was Major anions of all groundwater samples were determined by ion
directly precipitated as Ag2S in a trap of 0.1 M AgNO3 solution, with sub- chromatography (883 ICplus), and major cations were determined by
sequent filtration, washing, and drying of the precipitate as described by flame atomic absorption spectroscopy (Agilent Technologies 200 Series
Böttcher et al. (2000). AA, 240 AA). Only samples with a charge balance b±5% were accepted
Dry weights of all solid phase contents were corrected for the salt (n = 21). The alkalinity (predominantly bicarbonate) was determined
contribution of pore waters, which remained attached to the sediments with a Merck MColortest™ Carbonate Hardness Test (precision
after freeze drying (Appendix A.3). Sedimentary sodium contents ~0.1 mmol/L), and dissolved sulfide concentrations were determined
(Seibert et al., 2018) were used for this purpose. Moreover, TS using a visocolor ECO Test and a PF-12 photometric device. DOC
268 S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283

concentrations were measured by high temperature catalytic oxidation calculating the mean De. Before the measurement of De values, pre-
using a Shimadzu TOC-VCPH. A more detailed overview of the tests were carried out at different preheat temperatures (160 to
hydrochemical analyses is presented in Seibert et al. (2018). The 260 °C, in 20 °C steps) to find the most appropriate thermal treat-
hydrochemical composition of each sample was subsequently used to ment for the OSL measurements (Appendix A.5).
carry out speciation calculations and determine mineral saturation indi- The radionuclide concentrations (K, Th, and U) of 50 g oven-dried
ces (e.g., SIcalcite and SIsiderite), employing the geochemical software tool (130 °C) sediment of each subsample allotted for gamma spectrometry
PHREEQC (Parkhurst and Appelo, 2013). were measured for at least 24 h in plastic containers, using high resolu-
For the determination of the δ34S signatures of dissolved sulfate, the tion gamma spectrometry. The measurements were conducted not ear-
precipitated BaSO4 was mixed with V2O5 in Sn cups and combusted in a lier than four weeks after filling the containers to allow adjustment of
Thermo Flash 2000 EA elemental analyzer that was connected to a the radon disequilibrium. The β-dose rate and the γ-dose rate were cal-
Thermo Finnigan MAT 253 gas mass spectrometer via a Thermo Elec- culated using the conversion factors of Guérin et al. (2011) and the grain
tron Conflo IV split interface. Sulfur isotope measurements at Leibniz size dependent β-attenuation of Aitken (1985). By applying the water
IOW had a precision of better than ±0.2‰. For δ34S analysis of dissolved content attenuation factor of Aitken (1985), the β- and γ-dose rates
sulfide and TRIS, the precipitated ZnS was converted to Ag2S. Isotope ra- were corrected.
tios were converted to the V-CDT scale, using international reference The water content for sample 3596 was set to 3 ± 3 wt%
calibration materials and following the procedure given by Mann et al. (Table A.5.1) according to the mean water content of representative
(2009). subsamples (n = 9, σ = 1) and a conservative estimate of the range.
DOM was pre-concentrated and desalinated according to Dittmar Only sample 3596 was taken from depth above the ground water
et al. (2008), by passing the acidified (pH = 2, HCl p.a.) subsamples table (1.22 mbgs in April 2016). All other samples were taken from
over Bond ELUT PPL solid phase extraction cartridges. After washing depth below the mean groundwater table and were, therefore, con-
with acidified ultrapure water, the DOM extracts were eluted with LC- sidered to have been constantly water saturated before sampling.
MS grade methanol. Samples were 1:1 (v:v) diluted in a methanol- According to water content measurements of artificially water satu-
water matrix to 7 ppm DOC concentrations and measured by rated sand samples (n = 10, σ = 1) with representative pore vol-
electrospray-ionization Fourier transform ion cyclotron resonance umes (26 to 38%) and bulk densities (1.3 to 1.5 g cm−3), the water
mass spectrometry (ESI-FT-ICR-MS) on a 15T Bruker solariX instru- content for all other samples was set to 19 ± 5 wt%, again applying
ment. Sample measurement, calibration, and compound classification a conservative estimate of the range (Table A.5.1). The sediment
(e.g., peptide, polyphenol, or fatty acid compounds) protocols are de- dose rate was then obtained by using the corrected β- and γ-dose
scribed by Seidel et al. (2014, 2015). For each sample, weighed sums rates.
of the compound classes were calculated to determine site-specific The cosmic dose rate was determined according to Prescott and
DOM fingerprints. Hutton (1994) and Prescott and Stephan (1982). The total dose rate
was calculated by adding the sediment dose rate and the cosmic dose
2.5. Optically stimulated luminescence (OSL) dating rate (Table A.5.1). The OSL age of each sample was finally determined
by dividing the mean De by the total dose rate.
All sample preparation was carried out under subdued red light.
Oven dried (50 °C) samples were dry-sieved to obtain sand of 150 to 3. Results
200 μm in diameter. The sand was treated with 10% HCl and diluted so-
dium oxalate solution (Na2C2O4) to destroy carbonates and disperse ag- 3.1. Geologic characterization and sediment ages
gregates. Organic material was removed through wet oxidation using
30% hydrogen peroxide (H2O2). Density separations were carried out The upper 1.24 m of the core consisted of aeolian medium to fine-
using sodium polytungstate (SPT, Na6[H2W12O40]) to obtain minerals grained sands with low contents of shell debris (b1% dw., Fig. 2a,
of different densities, i.e., potassium feldspar (ρ b2.58 g cm−3), plagio- Fig. A.6.1). Sediments at greater depth were of marine origin and
clase (2.58 g cm−3 b ρ b 2.62 g cm−3), quartz (2.62 g cm−3 b ρ contained higher shell contents. A layer considerably enriched in shell
b 2.7 g cm−3) and heavy minerals (ρ N 2.7 g cm−3). Quartz grains debris (up to 36% dw.) and consisting of medium to coarse-grained
were selected for luminescence measurement and treated with 40% sands with low contributions of fine sand was found at ~7.00 to
hydrofluoric acid (HF) for one hour to etch the α-irradiated outer sur- 10.00 mbgs, only interrupted by a layer of medium to fine-grained
face of the grains and to remove remaining feldspar. Subsequently, the sands at 7.24 to 7.88 mbgs. Silt and clay fractions were very low (b1%
quartz grains were treated with 20% HCl for one hour and dry-sieved dw.) in all core sediments (Fig. A.6.1).
again, using a 150 μm sieve. Ultimately, the quartz grains were depos- OSL ages of the sands ranged between 78 ± 11 and 459 ± 52 a
ited on stainless steel discs with a diameter of 6 mm (medium aliquots), (Fig. 2b, Table A.5.2), with ages increasing over depth. Sedimentation
which had been sprayed with silicone oil before deposition to fix the rates were calculated based on the obtained ages and the vertical distance
grains. between each sampling point (Fig. 2b). Upper aeolian (b0.51 mbgs) as
OSL measurements were conducted with an automated lumines- well as marine sediments at 1.55 to 2.43 mbgs and 6.38 to 7.35 mbgs
cence reader (Risø TL/OSL DA-20) equipped with blue light stimula- were deposited at lower rates (0.5 to 0.7 cm/a), while lower aeolian sed-
tion at 470 nm and a 7.5 mm Hoya U-340 detection filter. The single- iments at 0.51 to 1.55 mbgs were deposited faster (3.7 cm/a). The highest
aliquot regenerative-dose (SAR) protocol of Murray and Wintle sedimentation rate (14.6 cm/a) was estimated for marine sediments at
(2000, 2003) was applied for pretests and determination of the 2.43 and 6.38 mbgs, using the youngest (296 a) and oldest (323
equivalent dose (De). To minimize the relative contribution of the a) sediment ages from this time of rapid deposition.
medium and slow components of the OSL signal, Early Background
Subtraction (EBG) was applied (Ballarini et al., 2007). OSL signals be- 3.2. Geochemistry
tween 0 and 0.48 s were calculated and background integration
limits were set from 0.48 to 1.28 s. It was checked that the recycling TIC contents were low in the upper 7 m of the core but increased at
ratios and IR/blue OSL depletion ratios of all aliquots were ≤ 10% from greater depths, which is in line with the observed shell contents (Fig. 3a,
unity according to the range of acceptability suggested by Wintle and c). TOC contents were very low throughout the core (TOCmean ~0.03%
Murray (2006). A one-sided Grubbs test (Grubbs, 1950) was applied dw.) and showed only little variation, with highest TOC contents
for one time to detect outliers within the measured D e values. D e being detected in the aeolian sands and at the core bottom (Fig. 3b). Ex-
values detected as outliers (P b 0.01) were excluded when tracted reactive iron fractions were low (Fedith. mean: ~10.6 mmol/kg,
S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283 269

Fig. 2. a) Geologic characterization of core sediments. Sample locations for grain size and SEM-EDX analyses are indicated (No. 1 to 5, Appendix A.6). * denotes that medium to coarse-
grained sands at the core bottom contained low fractions of fine sand (compare Fig. A.6.1). b) Sediment ages based on OSL dating (dots) as well as calculated sedimentation rates [cm/
a]. Bars at each dot present the uncertainty associated with the determined ages (Table A.5.2). Sediment ages in depth N 7.4 mbgs could not be determined due to high contents of
shell debris.

FeHCl mean: ~8.5 mmol/kg) and remarkably constant over depth Intact framboidal and euhedral pyrite minerals were only detected
(Fig. 3d). The offset between both reactive iron fractions was uniform, in the silt and clay fraction (b63 μm) of sample 4 and 5, which were col-
neglecting two outliers of FeHCl in the aeolian sands. TRIS was present lected in the brackish part of the core (Fig. 4). The diameter of the
at low contents (b0.2 mmol/kg) in the upper sediments and higher con- framboids ranged between ~6 and 14 μm, and the euhedral pyrite had
tents (up to ~3.0 mmol/kg) at depths N7 mbgs, where groundwater con- a width of ~4 μm.
ditions were brackish (Fig. 3e). TS contents were generally higher than
TRIS contents and point towards extremely low TRIS contents in the 3.3. Hydrochemistry
sediments from 2 to 7 mbgs, where TRIS could not be detected. Stable
sulfur isotope signatures of TRIS were lower in brackish (b−25‰) com- TDS concentrations indicated fresh groundwater (TDS b1 g/L) in the
pared to the fresh (N − 20‰) sediments (Fig. 3f). upper two monitoring wells (DN-a, -b), while groundwater of the lower

Fig. 3. Depth profiles of a) TIC, b) TOC, c) shell, d) reactive iron (Fedith. and FeHCl), and e) sulfur (TRIS and TS) contents, and f) stable sulfur isotope signatures of TRIS [‰ vs. V-CDT].
Horizontal lines mark the groundwater table (upper solid line) and the approximated transition from fresh to brackish salinities (lower dashed line) in spring 2016, based on
hydrochemical sampling in April 2016. Shell contents presented in c) correspond to the sediment fraction N1 mm (shell fragments) relative to the entire sample weight, while all
other species (a, b, d, and e) were determined for the sediment fraction b1 mm only. ‘bDL’ denotes samples for which TRIS contents could not be determined, as no ZnS precipitation
was observed during the extraction procedure.
270 S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283

Fig. 4. Pyrite minerals detected in the brackish sediments using SEM-EDX analysis. a) and b) show framboidal pyrite minerals, found in the sediment fraction b63 μm at 7.75 mbgs. c) and
d) show an euhedral and a framboidal pyrite mineral, respectively, found in the sediment fraction b63 μm at 9.48 mbgs.

two wells (DN-c, -d) was brackish (TDS N1 g/L) in April 2016, approxi- Although dissolved sulfide concentrations were below the detection
mately three weeks after core sampling in March 2016 (Fig. 5a). The sa- limit of the field tests, quantitative precipitation of dissolved sulfides
linity stratification was steady in spring, summer, and autumn. as ZnS was successful for the two brackish groundwater samples col-
However, after storm floods reached monitoring site DN in winter lected in September 2016. The stable sulfur isotope signatures of dis-
2016/17, seawater infiltrated and resulted in increased TDS concentra- solved sulfide in these samples were −28.3 and −14.3‰ vs. V-CDT,
tions in the upper two monitoring wells in January 2017 (5 to 12 g/L, respectively (Fig. 5l).
Fig. 5a). TDS levels of groundwater collected during sampling in With regards to DOM compositions, the freshest groundwater sam-
March 2017 were lower again (b5 g/L). The pH of all groundwater sam- ples were characterized by elevated relative amounts of less-reactive,
ples was slightly alkaline (7.5 to 8.2, Fig. 5b), with only small seasonal aromatic compounds, such as polycyclic aromates and polyphenol com-
variations. Equilibrium with respect to calcite was noticed for all sam- pounds (not presented). In contrast, labile DOM compounds, such as
ples (Fig. 5c). fatty acids (Fig. 6a), peptide compounds (Fig. 6b), and aliphatic,
Redox potentials decreased over depth and showed seasonal varia- heteroatom-containing sugars (Fig. 6c), generally increased with in-
tions, particularly in the upper two wells (Fig. 5d). DOC was present at creasing electrical conductivities and were at their highest in the seawa-
moderate concentrations (~0.1 to 0.7 mmol/L) in all samples (Fig. 5e). ter reference sample. Despite this overall trend, it was recognized that
Bicarbonate concentrations increased over depth, except for March the fresh groundwater samples DN-a and DN-b were relatively enriched
2017 when concentrations in the two bottom wells were lower com- in labile DOM fractions, which especially accounts for sugars.
pared to the previous sampling campaigns (Fig. 5f). Dissolved oxygen
was present in near-surface groundwater, decreased over depth, and 4. Discussion
was absent in the lowest well (Fig. 5g). Similar trends were observed
for nitrate, which showed high and seasonally varying concentrations 4.1. Geomorphologic changes, seasonal variations, and the impact on
of up to ~380 μmol/L in near-surface groundwater (Fig. 5h). Dissolved groundwater salinities
iron was detected in the lowest well at varying concentrations (b8 to
24 μmol/L, Fig. 5i) and one sample collected at the uppermost well in The upper 7.4 m of the investigated sediments were rapidly depos-
March 2017 (~10 μmol/L). Sulfate concentrations correlate with salinity ited, mainly under marine conditions within the previous ~500 years.
and were, thus, higher in the brackish compared to the freshwater zone Ages of sediments at N7.4 mbgs could not be determined but, based
(Fig. 5j). Infiltrating seawater increased sulfate concentrations in the on the petrography, it is assumed that they are not significantly older
former freshwater zone in January and March 2017. Concentrations of than ~900 a, applying a sedimentation rate of 0.7 cm/a as calculated
dissolved sulfide were always below the detection limit (b2 μmol/L, for the layer above. Aeolian sands at the top are younger than
not presented). ~80 years, which is in line with Röper et al. (2013) who reported that
Stable sulfur isotope signatures of dissolved sulfate ranged between the formation of dunes at the Ostplate started in the 1940's. The esti-
+19.5 and +21.7‰ vs. V-CDT in all brackish samples (Fig. 5k). Dis- mated sedimentation rates of ~0.5 to 14.6 cm/a at the Ostplate are
solved sulfate in freshwater collected at the upper two wells in Septem- very high compared to other marine locations, where sedimentation
ber 2016 was relatively depleted in 34S (+2.7 to +9.5‰ vs. V-CDT). rates typically are lower than 1 cm/a (e.g., Silverberg et al., 1986; van
S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283 271

Fig. 5. Results of hydrochemical sampling at monitoring site DN. a) TDS, b) pH, c) SICalcite, d) redox potential, e) DOC, f) bicarbonate, g) dissolved oxygen, h) nitrate, i) dissolved iron,
and j) dissolved sulfate concentrations, and k) δ34S-SO2− 34
4 and l) δ S-H2S signatures [‰ vs. V-CDT] are presented. The groundwater table (full horizontal lines) and the approximated
transition from fresh to brackish groundwater (dashed horizontal lines) are indicated, based on hydrochemical sampling in April 2016. ‘bDL’ denotes the detection limits for nitrate
(16 μmol/L) and dissolved iron (8 μmol/L), respectively.

Weering et al., 1987; Middelburg and de Lange, 1989; Calvert et al., late autumn, when no storm surges occur at the Ostplate, salinities
1991; Stein et al., 1994). quickly return to fresh conditions in the upper sediments due to infil-
While the study site remained fully marine until ~80 years ago, sub- trating precipitation and lateral groundwater flow from the freshwater
sequent dune growth inhibited the infiltration of seawater during tidal lens (Holt et al., 2017). Therefore, dune growth accompanied by the for-
cycles and most flood events during the year and allowed for the forma- mation of freshwater lenses together with the observed seasonal pat-
tion of freshwater lenses (Röper et al., 2013; Holt et al., 2017), resulting terns, including storm surges and periods of freshening, result in
in fresh to brackish groundwater salinities today. However, the study highly varying groundwater salinities at the core sampling site (Holt
site is occasionally impacted by storm surges in the winter season, tem- et al., in preparation). Changing salinities are further linked to the
porarily leading to increased groundwater salinities. During spring until biogeochemical cycles, either directly (e.g., changes of solute

Fig. 6. Relative concentrations of DOM for a) saturated fatty acids (CHOX), b) peptides, and c) sugars (CHOX) in groundwater samples from the monitoring sites DN (well DN-a, -b, -c, -d),
DS (well DS-b, -c), and B (well B-a, -b, -c), respectively, and a seawater sample collected in September 2016.
272 S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283

concentrations, such as dissolved sulfate) or indirectly (e.g., change of Both analyzed sedimentary reactive iron fractions are coupled to
the redox conditions and primary productivity), as stated by Böttcher each other and demonstrate common iron sources. A major portion of
and Lepland (2000). reactive iron in the core sediments can be linked to highly reactive fer-
rihydrite or lepidocrocite, which are extracted with the HCl method
(Raiswell et al., 1994). The offset between the Fedith. and FeHCl fractions
4.2. Characterization of carbon, sulfur, and iron sources (~2 to 4 mmol/kg) corresponds to a contribution from less reactive iron
(oxyhdr)oxides, e.g., goethite and hematite, which are quantitatively
Shell debris, constituting of low Mg-calcite and aragonite (Seibert extracted with the citrate-dithionite method (Canfield, 1989). Two
et al., 2018), are the main source for the TIC pool, based on correlated FeHCl outliers at the core top (Fig. 3d) may indicate a high local contribu-
TIC and shell contents (Fig. A.7.1a). The fact that TIC and TOC contents tion of silicate iron. Furthermore, PHREEQC speciation calculations for
do not correlate (Fig. A.7.1b) demonstrates the presence of mainly groundwater samples containing dissolved iron indicated that solutions
reworked shell fragments (Böttcher et al., 2004), i.e., organic matter as- were close to saturation/supersaturated with respect to siderite
sociated with shell debris was removed before or quickly degraded (SIsiderite ~0.1 to 0.5), which shows that iron carbonates possibly present
upon deposition. The observed TOC contents in the permeable sands additional iron sources. In general, reactive iron contents in the perme-
at the Ostplate are extremely low compared to other marine locations able sands are extremely low and comparable to other (organic-poor)
(e.g., Oenema, 1990; Lyons and Berner, 1992; Bottrell et al., 2009; permeable sands (e.g., Jakobsen and Postma, 1999; Morse, 1999;
Sheng et al., 2015) and even among the lowest that were reported for Böttcher et al., 2004).
similar organic-poor, permeable sands (e.g., Böttcher et al., 1998,
2004; Jakobsen and Postma, 1999). Besides sedimentary organic matter 4.3. Degradation of organic matter
(SOM), (labile) DOM presents a further carbon source in beach sedi-
ments subject to tidally- or storm-driven seawater inundation Dissolved oxygen and nitrate serve as the most prominent electron
(Chipman et al., 2010). acceptors for the oxidation of organic matter in near-surface groundwa-
Brackish groundwater samples have a Cl−/SO2− 4 ratio slightly higher ter, as both oxidants were rapidly consumed and absent at depth. Sim-
but similar to seawater at Spiekeroog Island (~18 mol*L−1/mol*L−1, ilar results for nitrate reduction were obtained by Allen et al. (2014),
Fig. 7), indicating that dissolved sulfate in brackish samples is of who investigated fresh groundwater below a natural dune system in
marine origin. Cl−/SO2− 4 ratios of freshwater samples are lower (~7 to SW England. Seasonal variations of nitrate concentrations may be ex-
12 mol ∗ L−1/mol ∗ L−1) and resemble the Cl−/SO2− 4 ratio of local precip- plained by varying nutrient loads linked to the activity of migrating
itation (~12 mol ∗ L−1/mol ∗ L−1). Stable sulfur isotope signatures con- birds, the oxidation of ammonia liberated during the degradation of or-
firm the previous considerations, because dissolved sulfate in all ganic matter, or contributions from atmospheric deposition and N2 fix-
brackish samples had a δ34S signature close to +21‰ vs. V-CDT, ation (Seibert et al., 2018).
which is typical for North Sea seawater sulfate (Böttcher et al., 2007), Reductive dissolution of iron (oxyhydr)oxides was inferred for the
and the two freshwater samples had δ34S-SO2− 4 values of +9.5 deepest monitoring well, where dissolved iron was detected. Dissolved
and +2.7‰ vs. V-CDT, respectively, which is similar to the typical iron observed in the uppermost sampling well in March 2017 was unex-
δ34 S-SO 2−
4 signatures in precipitation in North-West Germany pected as sampled groundwater contained significant concentrations of
(δ34S-SO2− 4 about +8 ± 2‰ vs. V-CDT; Reckhardt et al., 2017). Stable dissolved oxygen and nitrate, and this finding could be explained by the
sulfur isotope signatures of dissolved sulfate in coastal precipitation, input of labile marine organic matter during storm events in winter
which are lighter than the typical signature of seawater sulfate, are 2016/17 and the presence of reducing microniches. Such redox niches
an indication for non-marine sulfate sources, predominantly were, for instance, reported by Jørgensen (1977a) for sulfate reduction
reflecting emission and deposition of sulfur compounds from (indus- in marine sediments with oxidizing conditions in the surroundings.
trial) combustion processes (e.g., Caron et al., 1986; Novák et al., Maximum dissolved iron concentrations were up to two orders of mag-
2000) or, although presumably of minor importance, sulfur gas nitude lower compared to other organic-rich locations, where iron sul-
emissions from intertidal sediments (Kristensen et al., 2000). fide formation was observed (e.g., Oenema, 1990; Neumann et al., 2005;
Bottrell et al., 2009; Morgan et al., 2012). The low concentrations of dis-
solved iron in the investigated groundwater may further result from
siderite precipitation (e.g., Magaritz and Luzier, 1985), which is corrob-
orated by PHREEQC speciation calculations for selected groundwater
samples (SIsiderite ~0.1 to 0.5).
Microbial sulfate reduction could not be directly inferred from mea-
sured dissolved sulfide concentrations (b2 μmol/L), whereas an accu-
mulation of dissolved sulfide is often observed for pore waters of iron
sulfide-forming sediments (e.g., Lyons and Berner, 1992; Böttcher
et al., 1998; Morgan et al., 2012). However, the bulk precipitation of
ZnS was successful and calculated Cl−/SO2− 4 ratios of brackish samples
were similar to but slightly higher than the average ratio in seawater
(Fig. 7), which points towards ongoing microbial sulfate reduction
and, consequently, some sulfate consumption. These results are in line
with Magaritz and Luzier (1985) who observed non-conservative be-
havior of sulfate in the seawater-freshwater mixing zone of a coastal
dune aquifer at Coo Bay, Oregon, which was explained by microbial sul-
fate reduction.
Furthermore, a slight increase of δ34S-SO2− 4 values with increasing

Cl /SO2− 4 ratios was noticed for most brackish samples (Fig. 8b), and
Fig. 7. Cl−/SO2−
4 ratios of collected groundwater samples. The two vertical grey lines such an enrichment of 34S in the residual sulfate fraction is typically
indicate mean Cl−/SO2− 4 ratios of local seawater and precipitation, respectively,
according to Seibert et al. (2018). Horizontal lines mark the groundwater table (solid
found in natural closed-systems under the influence of sulfate reduction
line) and the approximated transition from fresh to brackish salinities (dashed line), (e.g., Hartmann and Nielsen, 2012). The enrichment of 34S in dissolved
based on hydrochemical sampling in April 2016. sulfate at the sampling site is, however, very low compared to other
S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283 273

Fig. 8. a) Cl−/SO2− 34 2−
4 ratios vs. δ S-SO4 signatures of collected groundwater samples. P and S (triangles) mark the composition of precipitation and seawater endmembers, respectively.
The dotted line presents the mixing line between both endmembers. The color scale of the dots indicates the salinity, with EC b1.5 mS/cm corresponding to freshwater. b) Cl−/SO2−
4 ratios
vs. δ34S-SO2−4 signatures of brackish groundwater samples.

terrestrial/marine locations (e.g., Böttcher et al., 1998; Massmann et al., (intertidal) sediments (e.g., Arnosti et al., 1998; Kristensen et al.,
2003; Neumann et al., 2005; Bottrell et al., 2009; Antler et al., 2013; 2000; Böttcher et al., 2004; Westrich and Berner, 1988).
Peketi et al., 2015), indicating extremely low sulfate reduction rates The fact that sulfate reduction was observed in the sediments was
and essentially open-system conditions (Fig. A.8.1). Pore water sulfate surprising, as reactive iron (oxyhdr)oxides (e.g., ferrihydrite and
replenishment at the study site, leading towards essentially open- lepidocrocite) were present. These are commonly expected to be uti-
system conditions with respect to sulfate, is caused by the seasonal in- lized as electron acceptors prior to the microbial reduction of sulfate
undation events, which further result in a mixing of fresh groundwater (e.g., Chapelle and Lovley, 1992). Results of this study are, therefore, in-
with sulfate-rich and oxygenated seawater. dicative for the concurrent reduction of iron (oxyhdr)oxides and sulfate,
Sulfate reduction rates (SRR) in the core sediments were estimated i.e., the presence of overlapping redox zones, as reported by Postma and
under the assumption that all produced dissolved sulfide reacted to Jakobsen (1996) for the Rømø aquifer, Denmark. Our findings may even
form pyrite: confirm recent incubation experiments with natural sediments by
Hansel et al. (2015), who showed that microbial sulfate reduction is ca-
pable of outcompeting poorly crystalline iron (oxyhydr)oxides as termi-
SRR ¼ Cpyrite  ρbulk =Δt ð1Þ
nal electron acceptors even at low sulfate levels.
The observed discrimination between δ34S-SO2− 34
4 and δ S-H2S (~35
with Cpyrite, corresponding to the pyrite content which was formed dur- to 49‰) is typical for sandy marine sediments in the region
ing a certain period of time, Δt, in sediments with the bulk density ρbulk. (e.g., intertidal sediments in the Spiekeroog Area: ~32 to 42‰,
Applying the maximum detected pyrite (TRIS) content of ~3 mmol/kg, a Böttcher et al., 1998; intertidal sediments at Westerhever: ~33 to
bulk density of 1.42 kg/L, that is typical for dune sands at the Ostplate 46‰, Böttcher et al., 2004). As experiments from pure cultures sug-
(Seibert et al., 2018), and a reaction time of 500 to 900 years for the gested a maximum isotope discrimination of ~46‰ during microbial
oldest sediments, the net SRR corresponds to ~0.01 to 0.02 nmol sulfate reduction (e.g., Chambers and Trudinger, 1979), higher values
∗ cm−3 ∗ d−1. This rate presents a minimum rate, as the majority of dis- were often attributed to recycling of sulfur species, including re-
solved sulfide usually undergoes re-oxidation in diffusion-driven ma- oxidation of sulfide followed by disproportionation of sulfur intermedi-
rine sediments (e.g., Thamdrup et al., 1994), and a maximum gross ates (e.g., Canfield and Thamdrup, 1994; Böttcher et al., 2005). More re-
SRR of ~0.2 to 0.4 nmol ∗ cm−3 ∗ d−1 would be obtained if it is assumed cently, however, Canfield et al. (2010) and Sim et al. (2011) could show
that ~95% of dissolved sulfide was re-oxidized (Jørgensen, 1982). SRRs that such large isotope fractionation does not necessarily require dis-
at the study site likely range between the calculated minimum and proportionation and can be achieved by dissimilatory sulfate reduction
maximum rates, as the present system is dominated by advective alone. In the absence of dissolved oxidizing agents (i.e., dissolved oxy-
groundwater flow, resulting in less re-oxidation compared to gen and nitrate), the observed isotope discrimination was assigned to
diffusion-driven marine sediments. Our findings are in line with microbial sulfate reduction and low process rates (Habicht and
Jakobsen and Postma (1999) who investigated a similar organic-poor Canfield, 1997). The overall sulfur isotope discrimination may further
coastal aquifer (Rømø Island, Denmark) and obtained maximum SRR be superimposed by some effect of the disproportionation of sulfur in-
of ~0.1 to 14 nmol ∗ cm−3 ∗ d−1, which were directly measured. SRR termediates upon re-oxidation with iron (oxyhydr)oxides.
in fully (brackish) marine sediments usually are considerably higher
and range between b10 and ~500 nmol ∗ cm−3 ∗ d−1 (e.g., Jørgensen, 4.4. Pyrite formation/oxidation and limiting factors
1977b; Hines and Lyons, 1982; Canfield et al., 1993b; Kristensen et al.,
2000; Böttcher et al., 2004; Bottrell et al., 2009) but may even reach Authigenic pyrite formation is inferred for the lower, brackish sedi-
values of up to 7000 nmol ∗ cm−3 ∗ d−1 (Habicht and Canfield, 1997). ments, where TRIS contents were elevated and sulfate reduction was
The very low SRR at the study site are attributed to the low content/re- observed. Furthermore, intact framboidal and euhedral pyrites were
activity of the present organic matter, as proposed by Jakobsen and present, which confirms the in-situ formation of pyrite (Wilkin et al.,
Postma (1999) for the Rømø aquifer and Kristensen et al. (2000) for in- 1996). Further evidence of the in-situ formation of pyrite is the match
tertidal sediments at Konigshafen. Furthermore, the relatively low in- between stable sulfur isotope signatures of dissolved sulfide and sedi-
situ temperatures, averaging ~10 °C, inhibit enhanced microbial activity. mentary TRIS, demonstrating that locally produced dissolved sulfide is
A general increase of SRR with increasing temperatures was reported immobilized due to pyrite precipitation. The important role of pyrite
for marine studies investigating seasonal SRR in near-surface formation at the anoxic interfaces of fresh and brackish/saline
274 S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283

groundwaters was previously noticed by Magaritz and Luzier (1985), (+2.7‰ vs. V-CDT) and precipitation (+8‰ vs. V-CDT), the stable sul-
who reported pyrite formation beneath the freshwater-seawater fur isotope signature of the additional sulfate source is ~ −6‰ vs. V-
mixing zone of a similar sandy coastal aquifer. Pyrite contents in the in- CDT. TRIS in the freshwater zone bears a similar stable isotope signature,
vestigated sediments are among the lowest reported for (organic-poor) demonstrating that the additional sulfate potentially sources in pyrite
permeable sands (e.g., Morse, 1999; Böttcher et al., 2004; Jakobsen and and was released during oxidation with dissolved oxygen or nitrate ac-
Postma, 1999). Iron sulfide contents in anoxic, organic-rich sediments cording to the following overall reaction pathways (Appelo and Postma,
often are up to three orders of magnitude higher (e.g., Middelburg, 2005):
1991; Lyons and Berner, 1992; Neumann et al., 2005).
Very low TRIS contents in the upper, fresh sediments may present an 15 7
FeS2 þ O2 þ H2 O→FeðOHÞ3 þ 2SO4 2− þ 4Hþ ð3Þ
inert, non-authigenic background pyrite fraction, which was 4 2
transported to the site during sediment deposition. This is in line with
the detection of TRIS in the unsaturated zone sediments, where pyrite 15
5FeS2 þ 15NO3 − þ 5H2 O þ 4Hþ → N2 þ 5FeOOH þ 10SO4 2− þ 9Hþ ð4Þ
formation can be ruled out because of oxic conditions. A further expla- 2
nation for low TRIS contents in the freshwater zone could be the oxida-
tion of authigenic pyrite as a consequence of salinity/redox changes. Up to ~47% of dissolved oxygen or ~32% nitrate, respectively, would
Under fully marine conditions, the site was exclusively influenced by be consumed during the oxidation of pyrite in the freshwater zone with
marine organic matter with a high reactivity (e.g., Cowie et al., 1992; respect to Eqs. (3) and (4), assuming a pyrite derived sulfate input of
Aller and Blair, 2004; Fig. 6), which probably resulted in the zone of sul- 0.08 mmol/L and considering maximum observed dissolved oxygen
fate reduction being located closer to the ground surface. Redox bound- and nitrate concentrations of 0.32 mmol/L and 0.38 mmol/L, respec-
aries were likely shifted to greater depths upon infiltration of fresh tively, in the upper freshwater well. Although the presented concept
groundwater with less reactive terrestrial organic matter, potentially of TRIS oxidation as an explanation for additional sulfate in fresh
resulting in the oxidation of previously formed pyrite. Pyrite oxidation groundwater appears reasonable and fits the field observations well, it
could further hold as an explanation for an observed sulfate surplus in is understood that previous consideration may suffer from uncer-
fresh groundwater with a Cl−/SO2− 4 ratio of ~7 mol/L/mol/L, which is tainties. For instance, unknown TRIS reactivity, pore water velocities,
lower than the typical ratios in the precipitation and seawater flow fields, or recharge amounts of oxic seawater during storm events
endmembers, respectively (Fig. 8a). In addition, the low δ34S-SO2− 4 may have an additional effect on TRIS oxidation and pore water sulfate
value of this sample (+2.7‰ vs. V-CDT) marks that the additional sul- concentrations, respectively.
fate source must be isotopically lighter than dissolved sulfate in both Regarding limiting factors for pyrite formation, there was no signifi-
precipitation and seawater (Fig. 8a). A two-component mixing calcula- cant C/S-correlation between TOC and TRIS contents (Fig. 9a), which is
tion was carried out to characterize the additional sulfate source: usually observed for diagenetic pyrite formation under sulfate reducing
conditions (e.g., Lin and Morse, 1991; Lyons and Berner, 1992). Thus,
CFW  δ34 SFW ¼ Cprecip:  δ34 Sprecip: þ Csource  δ34 Ssource ð2Þ carbon oxidation must proceed via consumption of more favorable elec-
tron acceptors, such as oxygen and nitrate. Moreover, it is possible that
with C and δ34S being the sulfate concentrations and the stable sulfur pyrite formation is not strictly linked to the oxidation of SOM but a dif-
isotope signatures of the fresh groundwater sample (FW), precipitation ferent DOM source passing the permeable sands, in which case sedi-
(precip.), and the additional sulfate source (source), respectively. Using mentary C/S ratios would not be applicable to identify carbon
the measured sulfate concentration of the fresh groundwater sample limitation on pyrite formation. As labile DOM fractions correlate with
(0.20 mmol/L) and assuming a sulfate contribution from precipitation salinity, the seasonal inundation events at the core sampling site result
(0.12 mmol/L) according to the measured chloride concentration in the infiltration of considerable amounts of seawater and, thus, (la-
(1.47 mmol/L) and the average Cl−/SO2− 4 ratio of local precipitation bile) DOM. Therefore, it is likely that labile, seawater-derived DOM
(~12 mmol/L/mmol/L), an additional sulfate contribution of drives the redox cycles besides SOM. The availability of dissolved sul-
~0.08 mmol/L was calculated, which corresponds to ~67% of the esti- fate, usually limiting pyrite formation in sediments with fresh to brack-
mated sulfate input from precipitation. According to Eq. (2) and using ish salinities (Berner and Raiswell, 1984), can further be excluded as a
the measured δ34S-SO2− 4 values of the fresh groundwater sample constraint on pyrite formation, since estimated sulfate reduction rates

Fig. 9. a) C/S plot of sedimentary TOC and TRIS contents. The line indicates C/S ratios of 2.8, corresponding to “normal marine” sediments (Berner and Raiswell, 1983), and b) calculated
Degree of Pyritization (DOP). The groundwater table (full horizontal line) and the approximated transition from fresh to brackish groundwater in April 2016 (dashed horizontal line) are
indicated.
S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283 275

were low and no depletion of sulfate was observed. Finally, the Degree • Microbial sulfate reduction is observed despite the presence of reac-
of Pyritization (DOP; Berner, 1970; Raiswell et al., 1988) was calculated tive iron (oxyhydr)oxides. This points towards the concurrent reduc-
to investigate the effect of iron limitation on pyrite formation: tion of sulfate and iron or might even point towards the microbial
consumption of sulfate prior to iron (oxyhdr)oxides.
Fepyrite • The stable sulfur isotope discrimination between dissolved sulfate and
DOP ¼  ð5Þ
Fepyrite þ Fereactive reduced sulfide is typical for sandy sediments that are open systems
with respect to sulfate, such as intertidal sediments in the North Sea
region.
where a value approaching 1 would be obtained if reactive iron is read-
• Low content and reactivity of SOM are identified as the major limiting
ily transformed to pyrite, which is usually the case under (semi-)euxinic
factor for pyrite formation in the organic-poor, permeable sands. In
conditions (e.g., Lyons and Berner, 1992). Assigning Fedith. as Fereactive
addition, it was ascertained that labile DOM correlates with salinity,
and TRIS (FeS2) as Fepyrite, low DOP values were calculated (b0.2,
and, as seasonal inundation events occur, labile, seawater-derived
Fig. 9b). Therefore, the amount of reactive iron (oxyhdr)oxides is not
DOM presents a further important source of electron donors in the in-
limiting pyrite formation. We concluded that the low content of reactive
vestigated sediments.
SOM presents the major limiting factor for pyrite formation in the inves-
tigated permeable sands.

Acknowledgments
5. Conclusions

The authors wish to thank the Nationalparkverwaltung


The aim of this study was to investigate the biogeochemical cycles of
Niedersächsisches Wattenmeer (NLPV), in particular G. Millat and C.
carbon, sulfur, and iron, including the formation of pyrite, in young and
Schulz, for the possibility to conduct studies on the Ostplate. The field
rapidly-deposited, permeable sands at the land-sea transition zone.
work was further supported by L. Diehl, T. Fresenborg, F. Grünenbaum,
Such sediment environments are present in many coastal regions
T. Haene, A. Harms, C. H. Lünsdorf, H. Madaj, M. Biel Maeso, and S.
worldwide, yet, they didn't receive much scientific attention. For this
Maiwald. Laboratory work was supported by E. Gründken, C. Lehners,
purpose, a 10 m sediment core was collected at a dune base of the bar-
U. Kücks, S. Plewe, I. Schmiedinger, H. Simon, and I. Ulber. We wish to
rier island Spiekeroog, Germany, where groundwater salinities vary
thank the members of Section 3 at the Leibniz Institute for Applied Geo-
from fresh to brackish. A combination of geochemical, hydrochemical,
physics (LIAG) for support during collecting and preparing the core
and stable sulfur isotope analyses, as well as OSL dating was applied.
samples for OSL dating. We are further grateful to the
The main conclusions of this study are:
Microbiogeochemistry Group at the ICBM and Leibniz IOW for the sup-
• OSL dating allows for the age determination of very young sediments port during geochemical sample preparation and analyses. We thank S.
(b500 years) in high resolution. Fock, C. Winkelmann, and C. Heithecker of the Nationalparkhaus
• Groundwater salinities strongly affect the cycling of redox sensitive Wittbülten, Spiekeroog, for the provision of accommodation and the
elements. While authigenic pyrite formation was observed in the support during field work. We highly acknowledge the constructive
brackish, anoxic sediments, pyrite oxidation was inferred for the comments of Simon Bottrell and one anonymous reviewer, and we ap-
fresh, oxic parts. preciate the excellent manuscript handling of the editor J.V. Cruz. We
• Estimated microbial sulfate reduction rates are extremely low, wish to thank the German Research Foundation for project funding
resulting in both low dissolved sulfide concentrations and pyrite con- (DFG project number MA 3274/6-1 and GI 171/25-1). This publication
tents. Comparable rates were only reported for other highly perme- is dedicated to our very supportive colleague Gerald Millat, who sadly
able, organic-poor sands. passed away in 2017.

Appendix A

A.1. Total Carbon (TC) measurements

Fig. A.1.1. Replicate measurements of total carbon for selected samples (n = 4). The deviation between the replicate measurements was b5% on average.
276 S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283

A.2. Fedith. measurements

Fig. A.2.1. Replicate measurements of Fedith.

A.3. Dry weight and TS corrections

The presented geochemical data was corrected for the salt content, which remained attached to the sediments after freeze-drying. For this pur-
pose, sodium contents [g/kg] were used (Seibert et al., 2018) and multiplied by a factor of 3.3, corresponding to the average ratio of Total Dissolved
Solids (TDS, i.e., salt) [mg/L] and Na [mg/L] in seawater at Spiekeroog Island (Fig. A.3.1). In a second step, all determined solid phase contents were
corrected using the following factor:
mdry
f salt ¼ 
mdry −msalt

where mdry and msalt correspond to the sediment dry weight and the dry weight of salt, respectively. Results show that the remaining amount of salt
(Fig. A.3.1) is very low in comparison to the sediment dry weights, ranging from ~22 to 38 g for all samples.

Fig. A.3.1. Measured sodium contents (Seibert et al., 2018) and estimated salt contents. Salt contents were calculated assuming an average ratio of 3.3 for TDS [mg/L] and Na [mg/L] in
seawater at Spiekeroog Island.
S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283 277

Moreover, measured TS contents were corrected for the contribution of pore water sulfate. Sodium contents (Fig. A.3.1) as well as interpo-
lated Na+/SO2−
4 ratios from groundwater sampling in April 2016 (Fig. A.3.2d) were used for this purpose. The calculated contribution of sulfate
on average corresponds to ~30% of the uncorrected TS content (Fig. A.3.2e).

Fig. A.3.2. a) TDS, b) sodium, and c) sulfate concentrations, and d) Na+/SO2− 4 ratios of groundwater samples collected in April 2016 (dots), and e) TS (full dots) and sulfate contents (open
circles) of the core sediments. Lines in a) to d) present linear interpolations between the sampling points. For the unsaturated zone, hydrochemical analyses of suction cup samples (n = 9)
were used. The suction cups were installed in the unsaturated zone at monitoring site DS in the direct vicinity of the core sampling site at the Ostplate (Fig. A.4.1). Sulfate contents were
obtained by dividing measured sodium contents (Seibert et al., 2018, Fig. A.3.1) by the interpolated pore water Na+/SO2− 4 ratio of the corresponding depth. The groundwater table (full
horizontal lines) and the approximated transition from fresh to brackish groundwater (dashed horizontal lines) are indicated, based on hydrochemical sampling in April 2016.

A.4. Dissolved organic matter (DOM) sampling

Samples for the characterization of DOM were collected at monitoring site DN in September 2016. Furthermore, groundwater samples were ob-
tained from 5 additional wells located at the monitoring site DS and B at the Ostplate, respectively (Fig. A.4.1). Moreover, a seawater sample was col-
lected for DOM analyses.
Two monitoring wells (DS-b, DS-c) are located at monitoring site DS, where the ground surface elevation is ~4.2 masl (Fig. A.4.1). Wells
have a screen length of 1 m and are screened in 4.7 to 5.7 mbgs (DS-b) and 6.7 to 7.7 mbgs (DS-c). Storm events usually do not reach site
DS, and groundwater electrical conductivities (EC) varied from 1.4 to 2.0 mS/cm (DS-b) and 7.6 to 8.5 mS/cm (DS-c), based on bimonthly mon-
itoring in 2016 to 2017.
Three monitoring wells (B-a, B-b, B-c) are located at monitoring site B at the northern beach of the Ostplate (Fig. A.4.1). The ground surface ele-
vation is ~1.8 masl, and, therefore, site B is regularly flooded during storm events. Wells have a screen length of 1 m and are screened in 0.5 to
1.5 mbgs (B-a), 6.6 to 7.6 mbgs (B-b), and 14.0 to 15.0 mbgs (B-c). Groundwater ECs varied from 30.5 to 36.2 mS/cm (B-a), 28.6 to 30.6 mS/cm
(B-b), and 26.2 to 33.7 mS/cm (B-c), based on bimonthly monitoring in 2016 to 2017.
Seawater for DOM analysis was collected at the Ostplate in September 2016 (Fig. A.4.1), and seawater ECs varied from 44.7 to 48.2 mS/cm
for sampling campaigns in 2016 to 2017. A more detailed description of the hydrochemistry is not presented in this study, as groundwater
samples from the monitoring wells at site DS and B only aid to investigate the characteristics and compositions of DOM at the Ostplate as a
function of salinity.
278 S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283

Fig. A.4.1. Locations of DOM sampling in September 2016. DN marks the core sampling site, including the groundwater monitoring wells (GMW) DN-a, DN-b, DN-c, and DN-d. DS and B
present additional monitoring sites, with the GMW DS-b, DS-c, and B-a, B-b, and B-c at the southern side of the dune ridge and the northern beach of the Ostplate, respectively. Suction
cups, used for the hydrochemical analysis of soil water, are located at site DS. Furthermore, the seawater sampling location is indicated. Seawater was collected during high tide. Yellow-
hatched areas present dune areas with ground surface elevations N3 masl, according to Röper et al. (2013). (For interpretation of the references to color in this figure legend, the reader is
referred to the web version of this article.)

A.5. Optically stimulated luminescence (OSL) dating

Prior to the De determination, pretests were performed on aliquots of one sample (3594) to find the most appropriate thermal treatment for
OSL measurements (Fig. A.5.1). This included a preheat plateau test, a thermal transfer test (for bleached aliquots), and a dose recovery test at
different preheat temperatures for 10 s and a fixed cut heat temperature of 160 °C. The preheat plateau test (Fig. A.5.1-A) showed a plateau
between 160 and 180 °C, from which the preheat temperature for De measurements should be selected, given the dose recovery ratio at
this temperature is within the acceptable range of ±10% from unity (Wintle and Murray, 2006). At the preheat temperature of 180 °C, the
recycling ratio was close to unity within the acceptable range and no contribution of thermal transferred OSL signals was observed
(Fig. A.5.1-B and -C). For this reason, the De measurements (16 aliquots per sample) and the determination of the dose recovery ratios (6 al-
iquots per sample) were conducted with a preheat temperature of 180 °C, using a cut heat temperature of 160 °C. An example of a typical OSL
decay curve and a dose response curve from an aliquot of sample 3594 is shown in Fig. A.5.2. The final results of OSL dating are presented in
Table A.5.2.
S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283 279

Fig. A.5.1. Pretests results of sample 3594. Equivalent dose (De) measured at preheat temperatures from 160 °C to 260 °C. A) Preheat plateau test, B) thermal transfer test, C) dose recovery
test. Single values are shown as dots and mean values (n = 4) as squares. Standard errors of the means are indicated by bars. During the preheat plateau test, one outlier (1.35 Gy, not
shown) was detected at a preheat temperature of 180 °C and excluded when calculating the mean De (n = 3).

Fig. A.5.2. Example of a typical OSL decay curve and a dose response curve (insert) of one aliquot of sample 3594. The regenerated signals are represented by squares and the
recycling point as a grey dot. Standard errors of the means are indicated by bars. The projection of the natural OSL signal (diamond) onto the dose response curve gave a De
of about 0.19 Gy.
280 S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283

Table A.5.1
Results of dose rate determinations.

Sample ID Depth Depth (mean) Grain size Water contenta Element concentration Cosmic dose rate Sediment dose rate Total dose rate

K U Th

[m] [m] [μm] [wt%] [%] [ppm] [ppm] [mGy a−1] [mGy a−1] [mGy a−1]

3596 0.44–0.58 0.51 150–200 3±3 0.39 ± 0.02 0.36 ± 0.02 0.92 ± 0.06 0.20 ± 0.02 0.51 ± 0.06 0.71 ± 0.06
3595 1.46–1.63 1.55 150–200 19 ± 5 0.52 ± 0.03 0.49 ± 0.03 1.20 ± 0.07 0.17 ± 0.02 0.57 ± 0.07 0.74 ± 0.07
3594 2.34–2.52 2.43 150–200 19 ± 5 0.46 ± 0.03 0.30 ± 0.02 0.72 ± 0.05 0.15 ± 0.02 0.47 ± 0.07 0.62 ± 0.07
3593 3.40–3.60 3.50 150–200 19 ± 5 0.45 ± 0.02 0.28 ± 0.02 0.68 ± 0.05 0.13 ± 0.01 0.45 ± 0.07 0.58 ± 0.07
3592 4.33–4.52 4.43 150–200 19 ± 5 0.69 ± 0.04 0.41 ± 0.03 1.12 ± 0.07 0.12 ± 0.01 0.69 ± 0.08 0.81 ± 0.08
3591 5.35–5.53 5.44 150–200 19 ± 5 0.72 ± 0.04 0.29 ± 0.02 0.82 ± 0.05 0.10 ± 0.01 0.68 ± 0.08 0.78 ± 0.08
3590 6.29–6.47 6.38 150–200 19 ± 5 0.75 ± 0.04 0.39 ± 0.03 0.85 ± 0.05 0.09 ± 0.01 0.72 ± 0.08 0.82 ± 0.08
3589 7.30–7.40 7.35 150–200 19 ± 5 0.85 ± 0.04 0.48 ± 0.03 1.27 ± 0.07 0.08 ± 0.01 0.84 ± 0.08 0.93 ± 0.08
a
Water content given in weight percentage of dry sediment.

Table A.5.2
Results of OSL dating. De is the equivalent dose.

Sample ID Depth Depth (mean) Dose recovery ratio Aliquotsa Mean De Ageb

[m] [m] [n] [mGy] [a]

3596 0.44–0.58 0.51 1.00 ± 0.01 15 55 ± 6 78 ± 11


3595 1.46–1.63 1.55 0.97 ± 0.02 15 79 ± 4 106 ± 11
3594 2.34–2.52 2.43 1.01 ± 0.01 16 185 ± 8 300 ± 35
3593 3.40–3.60 3.5 1.00 ± 0.01 16 186 ± 11 319 ± 41
3592 4.33–4.52 4.43 1.02 ± 0.01 15 262 ± 20 323 ± 39
3591 5.35–5.53 5.44 1.03 ± 0.02 15 243 ± 9 310 ± 32
3590 6.29–6.47 6.38 1.05 ± 0.02 15 242 ± 4 296 ± 29
3589 7.30–7.40 7.35 1.02 ± 0.02 16 425 ± 28 459 ± 52
a
Number of accepted aliquots for De determination.
b
OSL age calculated by dividing the mean De by the total dose rate in years (before 2017).

A.6. Grain size distribution

Fig. A.6.1. Grain size distribution of sediment samples collected at five different depths (compare Fig. 2a).
S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283 281

A.7. Carbon sources

Fig. A.7.1. a) TIC vs. shell contents including a linear regression (line), and b) TIC vs. TOC contents.

A.8. Degradation of organic matter and sulfate reduction

Fig. A.8.1. Cl− vs. SO2−


4 concentrations of collected groundwater samples. The line presents the mixing line between precipitation and seawater endmembers (Seibert et al., 2018). Colors
indicate the electrical conductivity (EC) of each sample.

References Berner, R.A., Raiswell, R., 1983. Burial of organic carbon and pyrite sulfur in sediments
over Phanerozoic time: a new theory. Geochim. Cosmochim. Acta 47 (5), 855–862.
Aitken, M.J., 1985. Thermoluminescence Dating. Academic Press, London. Berner, R.A., Raiswell, R., 1984. C/S method for distinguishing freshwater from marine
Allen, D., Darling, W.G., Williams, P.J., Stratford, C.J., Robins, N.S., 2014. Understanding the sedimentary rocks. Geology 12 (6), 365–368.
hydrochemical evolution of a coastal dune system in SW England using a multiple Blonder, B., Boyko, V., Turchyn, A.V., Antler, G., Sinichkin, U., Knossow, N., Klein, R.,
tracer technique. Appl. Geochem. 45, 94–104. Kamyshny Jr., A., 2017. Impact of aeolian dry deposition of reactive iron minerals
Aller, R.C., Blair, N.E., 2004. Early diagenetic remineralization of sedimentary organic C in on sulfur cycling in sediments of the Gulf of Aqaba. Front. Microbiol. 8, 1131.
the Gulf of Papua deltaic complex (Papua New Guinea): net loss of terrestrial C and Böttcher, M.E., Lepland, A., 2000. Biogeochemistry of sulfur in a sediment core from the
diagenetic fractionation of C isotopes. Geochim. Cosmochim. Acta 68 (8), 1815–1825. west-central Baltic Sea: evidence from stable isotopes and pyrite textures. J. Mar.
Al-Raei, A.M., Bosselmann, K., Böttcher, M.E., Hespenheide, B., Tauber, F., 2009. Seasonal Syst. 25 (3), 299–312.
dynamics of microbial sulfate reduction in temperate intertidal surface sediments: Böttcher, M.E., Oelschläger, B., Höpner, T., Brumsack, H.J., Rullkötter, J., 1998. Sulfate re-
controls by temperature and organic matter. Ocean Dyn. 59 (2), 351–370. duction related to the early diagenetic degradation of organic matter and “black
Antler, G., Turchyn, A.V., Rennie, V., Herut, B., Sivan, O., 2013. Coupled sulfur and oxygen spot” formation in tidal sandflats of the German Wadden Sea (southern North
isotope insight into bacterial sulfate reduction in the natural environment. Geochim. Sea): stable isotope (13C, 34S, 18O) and other geochemical results. Org. Geochem. 29
Cosmochim. Acta 118, 98–117. (5), 1517–1530.
Appelo, C.A.J., Postma, D., 2005. Geochemistry, Groundwater and Pollution. second ed. Böttcher, M.E., Hespenheide, B., Llobet-Brossa, E., Beardsley, C., Larsen, O., Schramm, A.,
CRC Press, Taylor & Francis Group, Amsterdam. Wieland, A., Böttcher, G., Berninger, U.-G., Amann, R., 2000. Biogeochemistry, stable
Arnosti, C., Jørgensen, B.B., Sagemann, J., Thamdrup, B., 1998. Temperature depen- isotope geochemistry, and microbial community structure of a temperate intertidal
dence of microbial degradation of organic matter in marine sediments: polysac- mudflat: an integrated study. Cont. Shelf Res. 20, 1749–1769.
charide hydrolysis, oxygen consumption, and sulfate reduction. Mar. Ecol. Prog. Böttcher, M.E., Hespenheide, B., Brumsack, H.J., Bosselmann, K., 2004. Stable isotope bio-
Ser. 59–70. geochemistry of the sulfur cycle in modern marine sediments: I. Seasonal dynamics
Ballarini, M., Wallinga, J., Wintle, A.G., Bos, A.J.J., 2007. A modified SAR protocol for optical in a temperate intertidal sandy surface sediment. Isot. Environ. Health Stud. 40 (4),
dating of individual grains from young quartz samples. Radiat. Meas. 42 (3), 360–369. 267–283.
Berner, R.A., 1970. Sedimentary pyrite formation. Am. J. Sci. 268 (1), 1–23. Böttcher, M.E., Thamdrup, B., Gehre, M., Theune, A., 2005. 34S/32S and 18O/16O fraction-
Berner, R.A., 1981. A new geochemical classification of sedimentary environments. ation during sulfur disproportionation by Desulfobulbus propionicus. Geomicrobiol J.
J. Sediment. Res. 51 (2). 22 (5), 219–226.
282 S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283

Böttcher, M.E., Brumsack, H.J., Dürselen, C.D., 2007. The isotopic composition of modern Jakobsen, R., Postma, D., 1999. Redox zoning, rates of sulfate reduction and interactions
seawater sulfate: I. Coastal waters with special regard to the North Sea. J. Mar. Syst. with Fe-reduction and methanogenesis in a shallow sandy aquifer, Rømø, Denmark.
67 (1), 73–82. Geochim. Cosmochim. Acta 63 (1), 137–151.
Bottrell, S.H., Mortimer, R.J., Davies, I.M., Martyn Harvey, S., Krom, M.D., 2009. Sulphur cy- Jørgensen, B.B., 1977a. Bacterial sulfate reduction within reduced microniches of oxidized
cling in organic-rich marine sediments from a Scottish fjord. Sedimentology 56 (4), marine sediments. Mar. Biol. 41 (1), 7–17.
1159–1173. Jørgensen, B.B., 1977b. The sulfur cycle of a coastal marine sediment (Limfjorden,
Burdige, D.J., 2007. Preservation of organic matter in marine sediments: controls, mecha- Denmark). Limnol. Oceanogr. 22 (5), 814–832.
nisms, and an imbalance in sediment organic carbon budgets. Chem. Rev. 107 (2), Jørgensen, B.B., 1982. Mineralization of organic matter in the sea bed - the role of sulphate
467–485. reduction. Nature 296 (5858), 643–645.
Calvert, S.E., Karlin, R.E., Toolin, L.J., Donahue, D.J., Southon, J.R., Vogel, J.S., 1991. Low Kaplan, I.R., Rittenberg, S.C., 1964. Microbiological fractionation of sulphur isotopes. Mi-
organic carbon accumulation rates in Black Sea sediments. Nature 350 (6320), crobiology 34 (2), 195–212.
692. Kraal, P., Burton, E.D., Bush, R.T., 2013. Iron monosulfide accumulation and pyrite forma-
Canfield, D.E., 1989. Reactive iron in marine sediments. Geochim. Cosmochim. Acta 53 tion in eutrophic estuarine sediments. Geochim. Cosmochim. Acta 122, 75–88.
(3), 619–632. Kristensen, E., Bodenbender, J., Jensen, M.H., Rennenberg, H., Jensen, K.M., 2000. Sulfur cy-
Canfield, D.E., Thamdrup, B., 1994. The production of 34S-depleted sulfide during bacte- cling of intertidal Wadden Sea sediments (Konigshafen, Island of Sylt, Germany): sul-
rial disproportionation of elemental sulfur. Science 266 (5193), 1973–1974. fate reduction and sulfur gas emission. J. Sea Res. 43 (2), 93–104.
Canfield, D.E., Thamdrup, B., Hansen, J.W., 1993a. The anaerobic degradation of organic Kunzmann, M., Bui, T.H., Crockford, P.W., Halverson, G.P., Scott, C., Lyons, T.W., Wing, B.A.,
matter in Danish coastal sediments: iron reduction, manganese reduction, and sulfate 2017. Bacterial sulfur disproportionation constrains timing of Neoproterozoic oxy-
reduction. Geochim. Cosmochim. Acta 57 (16), 3867–3883. genation. Geology 45 (3), 207–210.
Canfield, D.E., Jørgensen, B.B., Fossing, H., Glud, R., Gundersen, J., Ramsing, N.B., Thamdrup, LBEG (Landesamt für Bergbau, Energie und Geologie), 2018. Geologische Küstenkarte von
B., Hansen, J.W., Nielsen, L.P., Hall, P.O., 1993b. Pathways of organic carbon oxidation Niedersachsen 1:25000 (GHBK25) - Relief der Holozänbasis. NIBIS Kartenserver. URL.
in three continental margin sediments. Mar. Geol. 113 (1–2), 27–40. https://nibis.lbeg.de/cardomap3/, Accessed date: 4 April 2018.
Canfield, D.E., Farquhar, J., Zerkle, A.L., 2010. High isotope fractionations during sulfate re- Lin, S., Morse, J.W., 1991. Sulfate reduction and iron sulfide mineral formation in Gulf of
duction in a low-sulfate euxinic ocean analog. Geology 38 (5), 415–418. Mexico anoxic sediments. Am. J. Sci. 291 (1), 55–89.
Caron, F., Tessier, A., Kramer, J.R., Schwarcz, H.P., Rees, C.E., 1986. Sulfur and oxygen iso- Lin, Q., Wang, J., Algeo, T.J., Sun, F., Lin, R., 2016. Enhanced framboidal pyrite formation re-
topes of sulfate in precipitation and lakewater, Quebec, Canada. Appl. Geochem. 1 lated to anaerobic oxidation of methane in the sulfate-methane transition zone of the
(5), 601–606. northern South China Sea. Mar. Geol. 379, 100–108.
Chambers, L.A., Trudinger, P.A., 1979. Microbiological fractionation of stable sulfur iso- Lovley, D.R., 1991. Dissimilatory Fe (III) and Mn (IV) reduction. Microbiol. Rev. 55 (2),
topes: a review and critique. Geomicrobiol J. 1 (3), 249–293. 259–287.
Chapelle, F.H., Lovley, D.R., 1992. Competitive exclusion of sulfate reduction by Fe (lll)-re- Lovley, D.R., Goodwin, S., 1988. Hydrogen concentrations as an indicator of the predom-
ducing bacteria: a mechanism for producing discrete zones of high-iron ground inant terminal electron-accepting reactions in aquatic sediments. Geochim.
water. Groundwater 30 (1), 29–36. Cosmochim. Acta 52 (12), 2993–3003.
Chipman, L., Podgorski, D., Green, S., Kostka, J., Cooper, W., Huettel, M., 2010. Decomposi- Lovley, D.R., Phillips, E.J., 1987. Competitive mechanisms for inhibition of sulfate reduc-
tion of plankton-derived dissolved organic matter in permeable coastal sediments. tion and methane production in the zone of ferric iron reduction in sediments.
Limnol. Oceanogr. 55 (2), 857–871. Appl. Environ. Microbiol. 53 (11), 2636–2641.
Cline, J.D., 1969. Spectrophotometric determination of hydrogen sulfide in natural waters. Luther, G.W., 1991. Pyrite synthesis via polysulfide compounds. Geochim. Cosmochim.
Limnol. Oceanogr. 14 (3), 454–458. Acta 55 (10), 2839–2849.
Cowie, G.L., Hedges, J.I., Calvert, S.E., 1992. Sources and relative reactivities of amino acids, Luther, G.W., Church, T.M., 1988. Seasonal cycling of sulfur and iron in porewaters of a
neutral sugars, and lignin in an intermittently anoxic marine environment. Geochim. Delaware salt marsh. Mar. Chem. 23 (3–4), 295–309.
Cosmochim. Acta 56 (5), 1963–1978. Lyons, T.W., Berner, R.A., 1992. Carbon-sulfur-iron systematics of the uppermost deep-
Cypionka, H., Smock, A.M., Böttcher, M.E., 1998. A combined pathway of sulfur compound water sediments of the Black Sea. Chem. Geol. 99 (1–3), 1–27.
disproportionation in Desulfovibrio desulfuricans. FEMS Microbiol. Lett. 166 (2), Magaritz, M., Luzier, J.E., 1985. Water-rock interactions and seawater-freshwater mixing
181–186. effects in the coastal dunes aquifer, Coos Bay, Oregon. Geochim. Cosmochim. Acta
Dellwig, O., Böttcher, M.E., Lipinski, M., Brumsack, H.-J., 2002. Trace metals in Holocene 49 (12), 2515–2525.
coastal peats (NW Germany) and their relation to pyrite formation. Chem. Geol. Mann, J.L., Vocke, R.D., Kelly, W.R., 2009. Revised δ34S reference values for IAEA sulfur iso-
182, 423–442. tope reference materials S-2 and S-3. Rapid Commun. Mass Spectrom. 23,
Dittmar, T., Koch, B., Hertkorn, N., Kattner, G., 2008. A simple and efficient method for the 1116–1124.
solid-phase extraction of dissolved organic matter (SPE-DOM) from seawater. Marnette, E.C., van Breemen, N., Hordijk, K.A., Cappenberg, T.E., 1993. Pyrite formation in
Limnol. Oceanogr. Methods 6, 230–235. two freshwater systems in the Netherlands. Geochim. Cosmochim. Acta 57 (17),
Fossing, H., Jørgensen, B.B., 1989. Measurement of bacterial sulfate reduction in sedi- 4165–4177.
ments: evaluation of a single-step chromium reduction method. Biogeochemistry 8 Massmann, G., Tichomirowa, M., Merz, C., Pekdeger, A., 2003. Sulfide oxidation and sulfate
(3), 205–222. reduction in a shallow groundwater system (Oderbruch Aquifer, Germany). J. Hydrol.
Gagnon, C., Mucci, A., Pelletier, É., 1995. Anomalous accumulation of acid-volatile sul- 278 (1–4), 231–243.
phides (AVS) in a coastal marine sediment, Saguenay Fjord, Canada. Geochim. Merinero, R., Lunar, R., Somoza, L., Díaz-del-Río, V., Martínez-Frías, J., 2009. Nucleation,
Cosmochim. Acta 59 (13), 2663–2675. growth and oxidation of framboidal pyrite associated with hydrocarbon-derived sub-
Grubbs, F.E., 1950. Sample criteria for testing outlying observations. Ann. Math. Stat. 21 marine chimneys: lessons learned from the Gulf of Cadiz. Eur. J. Mineral. 21 (5),
(1), 27–58. 947–961.
Guérin, G., Mercier, N., Adamiec, G., 2011. Dose-rate conversion factors: update. Ancient Middelburg, J.J., 1991. Organic carbon, sulphur, and iron in recent semi-euxinic sediments
TL 29 (1), 5–8. of Kau Bay, Indonesia. Geochim. Cosmochim. Acta 55 (3), 815–828.
Habicht, K.S., Canfield, D.E., 1997. Sulfur isotope fractionation during bacterial sulfate Middelburg, J.J., de Lange, G.J., 1989. The isolation of Kau Bay during the last glaciation:
reduction in organic-rich sediments. Geochim. Cosmochim. Acta 61 (24), direct evidence from interstitial water chlorinity. Neth. J. Sea Res. 24 (4), 615–622.
5351–5361. Moore, J.K., Braucher, O., 2008. Sedimentary and mineral dust sources of dissolved iron to
Habicht, K.S., Canfield, D.E., 2001. Isotope fractionation by sulfate-reducing natural popu- the world ocean. Biogeosciences 5 (3), 631–656.
lations and the isotopic composition of sulfide in marine sediments. Geology 29 (6), Morgan, B., Burton, E.D., Rate, A.W., 2012. Iron monosulfide enrichment and the presence
555–558. of organosulfur in eutrophic estuarine sediments. Chem. Geol. 296, 119–130.
Habicht, K.S., Canfield, D.E., Rethmeier, J., 1998. Sulfur isotope fractionation during bacte- Morse, J.W., 1999. Sulfides in sandy sediments: new insights on the reactions responsible
rial reduction and disproportionation of thiosulfate and sulfite. Geochim. Cosmochim. for sedimentary pyrite formation. Aquat. Geochem. 5 (1), 75–85.
Acta 62 (15), 2585–2595. Morse, J.W., Thomson, H., Finneran, D.W., 2007. Factors controlling sulfide geochemistry
Hansel, C.M., Lentini, C.J., Tang, Y., Johnston, D.T., Wankel, S.D., Jardine, P.M., 2015. Domi- in sub-tropical estuarine and bay sediments. Aquat. Geochem. 13 (2), 143–156.
nance of sulfur-fueled iron oxide reduction in low-sulfate freshwater sediments. Murray, A.S., Wintle, A.G., 2000. Luminescence dating of quartz using an improved single-
ISME J. 9 (11), 2400–2412. aliquot regenerative-dose protocol. Radiat. Meas. 32 (1), 57–73.
Hartmann, M., Nielsen, H., 2012. δ34S values in recent sea sediments and their signifi- Murray, A.S., Wintle, A.G., 2003. The single aliquot regenerative dose protocol: potential
cance using several sediment profiles from the western Baltic Sea. Isot. Environ. for improvements in reliability. Radiat. Meas. 37 (4), 377–381.
Health Stud. 48 (1), 7–32. Nakai, N., Jensen, M.L., 1964. The kinetic isotope effect in the bacterial reduction and ox-
Hines, M.E., Lyons, W.B., 1982. Biogeochemistry of nearshore Bermuda sediments. I. Sul- idation of sulfur. Geochim. Cosmochim. Acta 28 (12), 1893–1912.
fate reduction rates and nutrient generation. Mar. Ecol. Prog. Ser. 87–94. Nealson, K.H., Myers, C.R., 1990. Iron reduction by bacteria: a potential role in the genesis
Holt, T., Seibert, S.L., Greskowiak, J., Freund, H., Massmann, G., 2017. Impact of storm tides of banded iron formations. Am. J. Sci. 290, 35–45.
and inundation frequency on water table salinity and vegetation on a juvenile barrier Neretin, L.N., Böttcher, M.E., Jørgensen, B.B., Volkov, I.I., Lüschen, H., Hilgenfeldt, K., 2004.
island. J. Hydrol. 554, 666–679. Pyritization processes and greigite formation in the advancing sulfidization front in
Holt, T., Greskowiak, J., Seibert, S.L., Massmann, G., 2018. Modelling the evolution of a the Upper Pleistocene sediments of the Black Sea. Geochim. Cosmochim. Acta 68
freshwater lens under highly dynamic conditions on a currently developing barrier (9), 2081–2093.
island on a decadal timescale (In preparation). Neumann, T., Rausch, N., Leipe, T., Dellwig, O., Berner, Z., Böttcher, M.E., 2005. Intense py-
Howarth, R.W., 1984. The ecological significance of sulfur in the energy dynamics of salt rite formation under low-sulfate conditions in the Achterwasser lagoon, SW Baltic
marsh and coastal marine sediments. Biogeochemistry 1 (1), 5–27. Sea. Geochim. Cosmochim. Acta 69 (14), 3619–3630.
Huerta-Diaz, M.A., Tessier, A., Carignan, R., 1998. Geochemistry of trace metals associated NLWKN (Niedersächsischer Landesbetrieb für Wasserwirtschaft, Küsten- und
with reduced sulfur in freshwater sediments. Appl. Geochem. 13 (2), 213–233. Naturschutz), 2014. Laser-Scan Elevation Model of 2014.
S.L. Seibert et al. / Science of the Total Environment 649 (2019) 264–283 283

Novák, M., Kirchner, J.W., Groscheová, H., Havel, M., Černý, J., Krejčí, R., Buzek, F., 2000. Seidel, M., Beck, M., Greskowiak, J., Riedel, T., Waska, H., Suryaputra, I.G.N.A., Schnetger, B.,
Sulfur isotope dynamics in two Central European watersheds affected by high atmo- Niggemann, J., Simon, M., Dittmar, T., 2015. Benthic-pelagic coupling of nutrients and
spheric deposition of SOx. Geochim. Cosmochim. Acta 64 (3), 367–383. dissolved organic matter composition in an intertidal sandy beach. Mar. Chem. 176,
Oenema, O., 1990. Pyrite accumulation in salt marshes in the Eastern Scheldt, southwest 150–163.
Netherlands. Biogeochemistry 9 (1), 75–98. Sheng, Y., Sun, Q., Shi, W., Bottrell, S., Mortimer, R., 2015. Geochemistry of reduced inor-
Parkhurst, D.L., Appelo, C.A.J., 2013. Description of input and examples for PHREEQC ver- ganic sulfur, reactive iron, and organic carbon in fluvial and marine surface sediment
sion 3: a computer program for speciation, batch-reaction, one-dimensional trans- in the Laizhou Bay region, China. Environ. Earth Sci. 74 (2), 1151–1160.
port, and inverse geochemical calculations (No. 6-A43). US Geological Survey. Silverberg, N., Nguyen, H., Delibrias, G., Koide, M., Sundby, B., Yokoyama, Y., Chesselet, R.,
Passier, H.F., Middelburg, J.J., de Lange, G.J., Böttcher, M.E., 1997. Pyrite contents, 1986. Radionuclide profiles, sedimentation-rates, and bioturbation in modern sedi-
microtextures, and sulfur isotopes in relation to formation of the youngest eastern ments of the Laurentian Trough, Gulf-of-St-Lawrence. Oceanol. Acta 9 (3), 285–290.
Mediterranean sapropel. Geology 25 (6), 519–522. Sim, M.S., Bosak, T., Ono, S., 2011. Large sulfur isotope fractionation does not require dis-
Peketi, A., Mazumdar, A., Joao, H.M., Patil, D.J., Usapkar, A., Dewangan, P., 2015. Coupled proportionation. Science 333 (6038), 74–77.
C\\S\ \Fe geochemistry in a rapidly accumulating marine sedimentary system: diage- Sindowski, K.H., 1968. Gliederungsmöglichkeiten im sandig ausgebildeten Küsten-
netic and depositional implications. Geochem. Geophys. Geosyst. 16 (9), 2865–2883. Holozän Ostfrieslands. Eiszeit. Gegenw. 19 (1), 209–218.
Postma, D., Jakobsen, R., 1996. Redox zonation: equilibrium constraints on the Fe (III)/ Sindowski, K.H., 1970. Erläuterungen zu Blatt Spiekeroog Nr. 2212. Niedersächsisches
SO4-reduction interface. Geochim. Cosmochim. Acta 60 (17), 3169–3175. Landesamt für Bodenforschung, Hannover.
Prescott, J.R., Hutton, J.T., 1994. Cosmic ray distribution to dose rates for luminescence and Stein, R., Schubert, C., Vogt, C., Fütterer, D., 1994. Stable isotope stratigraphy, sedimenta-
ESR dating: large depths and long-term variations. Radiat. Meas. 23 (2–3), 497–500. tion rates, and salinity changes in the Latest Pleistocene to Holocene eastern central
Prescott, J.R., Stephan, L.G., 1982. The contribution of cosmic radiation to the environmen- Arctic Ocean. Mar. Geol. 119 (3–4), 333–355.
tal dose for thermoluminescent dating: latitude, altitude and depth dependences. Stookey, L.L., 1970. Ferrozine – a new spectrophotometric reagent for iron. Anal. Chem. 42
Pact 6, 17–25. (7), 779–781.
Raiswell, R., Canfield, D.E., 2012. The iron biogeochemical cycle past and present. Streif, H., 1990. Das ostfriesische Küstengebiet. Nordsee, Inseln, Watten und Marschen;
Geochem. Perspect. 1 (1), 1–2. mit 10 Tabellen. Berlin-Stuttgart, Borntraeger (in German).
Raiswell, R., Buckley, F., Berner, R.A., Anderson, T.F., 1988. Degree of pyritization of iron as Tagliabue, A., Bopp, L., Aumont, O., 2009. Evaluating the importance of atmospheric and
a paleoenvironmental indicator of bottom-water oxygenation. J. Sediment. Res. 58 sedimentary iron sources to Southern Ocean biogeochemistry. Geophys. Res. Lett.
(5), 812–819. 36 (13).
Raiswell, R., Canfield, D.E., Berner, R.A., 1994. A comparison of iron extraction methods for Thamdrup, B., Fossing, H., Jørgensen, B.B., 1994. Manganese, iron and sulfur cycling in a
the determination of degree of pyritisation and the recognition of iron-limited pyrite coastal marine sediment, Aarhus Bay, Denmark. Geochim. Cosmochim. Acta 58
formation. Chem. Geol. 111 (1), 101–110. (23), 5115–5129.
Raiswell, R., Benning, L.G., Tranter, M., Tulaczyk, S., 2008. Bioavailable iron in the Southern Tronicke, J., 1997. Geophysikalische Erkundung von Süßwasservorkommen und
Ocean: the significance of the iceberg conveyor belt. Geochem. Trans. 9 (1), 7. Infiltrationszonen in Inselaquiferen. Diploma thesis. Institut für Geophysik der
Reckhardt, A., Beck, M., Greskowiak, J., Schnetger, B., Böttcher, M.E., Gehre, M., Brumsack, Westfälischen Wilhelms-Universität Münster.
H.-J., 2017. Cycling of redox-sensitive elements in a sandy estuary of the southern Van Weering, T.C., Berger, G.W., Kalf, J., 1987. Recent sediment accumulation in the Skag-
North Sea. Mar. Chem. 188, 6–17. errak, northeastern North Sea. Neth. J. Sea Res. 21 (3), 177–189.
Rickard, D., Luther, G.W., 1997. Kinetics of pyrite formation by the H2S oxidation of iron Westrich, J.T., Berner, R.A., 1988. The effect of temperature on rates of sulfate reduction in
(II) monosulfide in aqueous solutions between 25 and 125 C: the mechanism. marine sediments. Geomicrobiol J. 6 (2), 99–117.
Geochim. Cosmochim. Acta 61 (1), 135–147. Wilkin, R.T., Barnes, H.L., 1996. Pyrite formation by reactions of iron monosulfides with
Rickard, D., Morse, J.W., 2005. Acid volatile sulfide (AVS). Mar. Chem. 97 (3), 141–197. dissolved inorganic and organic sulfur species. Geochim. Cosmochim. Acta 60 (21),
Röper, T., Greskowiak, J., Freund, H., Massmann, G., 2013. Freshwater lens formation 4167–4179.
below juvenile dunes on a barrier island (Spiekeroog, Northwest Germany). Estuar. Wilkin, R.T., Barnes, H.L., Brantley, S.L., 1996. The size distribution of framboidal pyrite in
Coast. Shelf Sci. 121, 40–50. modern sediments: an indicator of redox conditions. Geochim. Cosmochim. Acta 60
Rudnicki, M.D., Elderfield, H., Spiro, B., 2001. Fractionation of sulfur isotopes during bacte- (20), 3897–3912.
rial sulfate reduction in deep ocean sediments at elevated temperatures. Geochim. Wintle, A.G., Murray, A.S., 2006. A review of quartz optically stimulated luminescence
Cosmochim. Acta 65 (5), 777–789. characteristics and their relevance in single-aliquot regeneration dating protocols.
Schoonen, M.A., 2004. Mechanisms of sedimentary pyrite formation. Geol. Soc. Am. Spec. Radiat. Meas. 41 (4), 369–391.
Pap. 379, 117–134. Wortmann, U.G., Bernasconi, S.M., Böttcher, M.E., 2001. Hypersulfidic deep biosphere in-
Seibert, S.L., Holt, T., Reckhardt, A., Ahrens, J., Pollmann, T., Giani, L., Waska, H., Böttcher, dicates extreme sulfur isotope fractionation during single-step microbial sulfate re-
M.E., Greskowiak, J., Massmann, G., 2018. Hydrochemical evolution of a freshwater duction. Geology 29 (7), 647–650.
lens below a barrier island (Spiekeroog, Germany): the role of carbonate mineral re- WSV (Wasser- und Schifffahrtsverwaltung des Bundes), 2016. High and Low Water Data
actions, cation exchange and redox processes. Appl. Geochem. 92, 196–208. at Tide Gauge Spiekeroog and Wangerooge 1995–2016. Provided by. Bundesanstalt
Seidel, M., Beck, M., Riedel, T., Waska, H., Suryaputra, I.G.N.A., Schnetger, B., Niggemann, J., für Gewässerkunde (BfG) (Unpublished results).
Simon, M., Dittmar, T., 2014. Biogeochemistry of dissolved organic matter in an an-
oxic intertidal creek bank. Geochim. Cosmochim. Acta 140, 418–434.

Vous aimerez peut-être aussi