You are on page 1of 14

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/323703351

Geochemical Effects of Millimolar Hydrogen Concentrations in Groundwater:


An Experimental Study in the Context of Subsurface Hydrogen Storage

Article in Environmental Science and Technology · March 2018


DOI: 10.1021/acs.est.7b05467

CITATIONS READS

19 542

5 authors, including:

Marton Berta Frank Dethlefsen


Christian-Albrechts-Universität zu Kiel Christian-Albrechts-Universität zu Kiel
14 PUBLICATIONS 41 CITATIONS 35 PUBLICATIONS 398 CITATIONS

SEE PROFILE SEE PROFILE

Markus Ebert
Christian-Albrechts-Universität zu Kiel
55 PUBLICATIONS 870 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Uncertainties of geochemical modelling by geochemical codes and thermodynamic databases View project

KORA TP3.1 Hannover Südstadt View project

All content following this page was uploaded by Marton Berta on 13 November 2018.

The user has requested enhancement of the downloaded file.


Article

Cite This: Environ. Sci. Technol. XXXX, XXX, XXX−XXX pubs.acs.org/est

Geochemical Effects of Millimolar Hydrogen Concentrations in


Groundwater: An Experimental Study in the Context of Subsurface
Hydrogen Storage
Marton Berta,* Frank Dethlefsen,* Markus Ebert, Dirk Schaf̈ er, and Andreas Dahmke
Department of Applied Geology, Aquatic Geochemistry and Hydrogeology, Institute of Geoscience, Kiel University,
Ludewig-Meyn-Straße 10, 24118 Kiel, Germany
*
S Supporting Information

ABSTRACT: Hydrogen storage in geological formations is


one of the most promising technologies for balancing major
fluctuations between energy supply from renewable energy
plants and energy demand of customers. If hydrogen gas is
stored in a porous medium or if it leaks into a shallow aquifer,
redox reactions can oxidize hydrogen and reduce electron
acceptors such as nitrate, FeIII and MnIV (hydro)oxides, sulfate,
and carbonate. These reactions are of key significance, because
they can cause unintentional losses in hydrogen stored in
porous media and they also can cause unwanted changes in the
composition of protected potable groundwater. To represent
an aquifer environment enclosing a hydrogen plume,
laboratory experiments using sediment-filled columns were
constructed and percolated by groundwater in equilibrium with high (2−15 bar) hydrogen partial pressures. Here, we show that
hydrogen is consumed rapidly in these experiments via sulfate reduction (18 ± 5 μM h−1) and acetate production (0.030 ± 0.006
h−1), while no methanogenesis took place. The observed reaction rates were independent from the partial pressure of hydrogen
and hydrogen consumption only stopped in supplemental microcosm experiments where salinity was increased above 35 g L−1.
The outcomes presented here are implemented for planning the sustainable use of the subsurface space within the ANGUS+
project.

■ INTRODUCTION
The energy production by wind turbines and solar panels
large quantities of gaseous hydrogen.2−9 Moreover, sustainable,
renewable-based energy systems will likely require a series of
fluctuates due to variable wind velocity and solar radiation, subsurface energy storage technologies simultaneously. Such
while the energy demand of consumers also varies creating a technologies include chemical (hydrogen, synthetic methane),
significant challenge for the stability of future energy supplies mechanical (compressed air), or heat energy storage, facing the
based mainly on renewable sources. One concept to address considerable amount of energy that needs to be stored and the
this issue is the “power to gas” approach, where the generation lack of other available storage technologies in the short or
of a gas with high energy content uses the excess electrical medium term fulfilling volume, cost, and response time
energy at times where the public energy demand is smaller than requirements.1
the produced energy. Large scale gas storage is a potentially Power-to-gas energy storage including geological hydrogen
necessary element in this concept, allowing the gas to be storage is expected to play a potentially deciding role in future
available for reuse when the demand for energy exceeds the energy networks, considering the following reasons. First, large-
supply. The overall difference between the amount of energy scale development of renewable energy sources became a
supplied by renewable energy sources and the amount of cornerstone of all national energy policies following the Paris
energy demanded usually varies in time, but this gap may reach Agreement,10 even if those actions are still unsatisfying
10 GWh−10 TWh, calculated as an example on a time scale of compared to the volume of development required to reach
days to months on a German regional scale power network.1 climate protection goals already agreed on internationally.11
Electrochemical generation, storage, and consumption of Second, consequently, estimations on for instance German
molecular hydrogen (in the following just “hydrogen”) has energy supplies project up to 100% renewable power
the potential to balance such fluctuating gaps.1,2 The volume of
hydrogen corresponding to the amount of energy to be Received: October 25, 2017
balanced in an energy network ranges between millions to Revised: March 10, 2018
billions of normal cubic meters.2 The subsurface may offer a Accepted: March 12, 2018
comparatively cost-efficient and safe option for storing such Published: March 12, 2018

© XXXX American Chemical Society A DOI: 10.1021/acs.est.7b05467


Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

Table 1. Hydrogen Oxidizing Reactions with Associated Gibbs Free Energy Yields under Standard Conditions and Neutral pH
and the Concentrations of Hydrogen Typical for Environments with Characteristic Metabolic Processes

reactions
ΔG0 [kJ·mol(H2)−1] H2 [nmol·L−1] references
aerobic hydrogen oxidation 2H2 + O2 → 2H2O −237 n.d. 73

hydrogen oxidation coupled to denitrification 5H2 + 2H+ + 2NO3− → N2 + 6H2O −224 0.03 25, 46, 73
producing N2

hydrogen oxidation coupled to nitrate reduction 4H2 + 2H+ + NO3− → NH4+ + 3H2O −150 <0.05 40, 73
producing NH3

hydrogen oxidation coupled to MnIV reduction H2 + MnO2 → Mn(OH)2 −163 <0.05 25, 46, 74

hydrogen oxidation coupled to FeIII reduction H2 + 2Fe(OH)3 → 2Fe(OH)2 + 2H2O −114 0.2 25, 46, 73

hydrogenotrophic sulfate reduction 4H2 + H+ + SO42− → HS− + 4H2O −57 1−2 25, 40, 46, 73

hydrogenotrophic methanogenesis 4H2 + H+ + HCO3− → CH4 + 3H2O −34 5−10 25, 40, 46, 73, 75

hydrogenotrophic acetogenesis 4H2 + H+ + 2HCO3− → CH3COO− + 4H2O −26 100< 25, 40, 75

generation by 2050.12 Third, if more than ca. 80% of the energy potential with increasing time or flow path length) because the
is produced by renewable sources in a power network, then the electron accepting processes coupled to hydrogen oxidation
inclusion of large-scale storage facilities, like subsurface with the higher energy yield (e.g., sulfate reduction) out-
hydrogen storage, is suggested to be a prerequisite for stable compete the others (e.g., acetogenesis) by keeping the
and profitable operation.1 Overview studies on technical aspects hydrogen concentrations below a threshold concentration
of subsurface hydrogen storage were published in the past few required for that process (Table 1).42,43 Among other
years13−16 as a part of an increasing research and industrial possibilities, such reactive environments may be numerically
interest, but field or experimental studies specifying potential simulated assuming a dual substrate limitation by hydrogen
environmental impacts or operational issues linked to storing concentration on the one hand and the concentration of sulfate
large amounts of pure hydrogen at elevated pressures in or total dissolved inorganic carbon species on the other.
geological media are missing so far. However, studies on Furthermore, not only the occurrence but also the kinetics of
hydrogen-containing gas mixtures stored in porous town gas the redox reactions is usually dependent on the concentration
reservoirs hint that hydrogenotrophic redox reactions poten- of dissolved hydrogen in those hydrogen-limited reactive
tially consume hydrogen at such conditions.14,17−21 These environments, and such processes are often modeled using
reactions may cause losses from the stored gas, and they may Monod kinetics.44−48
also change the composition of reservoir water. Furthermore, In situ remediation of contaminated aquifers may lead to
accidental leaking of hydrogen into shallow aquifers22 may elevated dissolved hydrogen concentration, e.g., when applying
result in conditions where an enlarged hydrogen partial zerovalent iron (ZVI)49−51 or hydrogen releasing com-
pressure (i.e., equal to the hydrostatic pressure in case of a pounds,52 in hydrogenotrophic denitrification systems,53−57 or
pure hydrogen gas phase) will cause a high dissolved hydrogen during the sulfidogenic treatment of wastewaters containing
concentration in the shallow groundwater, probably initiating dissolved metal ions.58,59 Increased dissolved hydrogen
typical redox reactions associated with hydrogen oxidation. concentration is thereby either an aim of the technology
Hydrogen-driven redox reactions are predominantly micro- (e.g., denitrification or sulfidogenic treatment) or a side-effect
biologically catalyzed23 and well-known from aquatic geo- (e.g., ZVI application for reductive dehalogenation). Usually,
systems, where hydrogen is the electron donor and O2, NO3−, those systems also show the establishment of redox sequences
FeIII and MnIV(hydr)oxides, SO42−, or dissolved CO2 are known from pristine aquifers, and the kinetics of the occurring
usually the terminal electron acceptors (Table 1). These redox redox reactions also depend on the hydrogen concentrations
reactions may produce NO2−, N2O, N2, NH4+, MnII, FeII, H2S, created by hydrogen-producing primary reactions.46 Further-
CH3COOH, or CH4, which might be released into the pore more, the hydrogen concentration found in such reactive
water or precipitated into various mineral phases. Reaction environments is mostly limited to a few hundred micromoles
products such as NO2−, H2S, or CH4 may have a negative effect per liter.
on the composition of the groundwater in terms of its usability In the context of the storage of pure gaseous hydrogen at
for other applications, e.g., drinking water production.24 pressures between 30 and 400 bar60−62 in the subsurface,
Moreover, groundwater itself is usually a protected natural dissolution of hydrogen may result in concentrations of several
good. In pristine aquifers, fermentative degradation of peptides, millimoles per liter groundwater due to the roughly linear
saccharides, and further organic substances produces hydro- correlation between increasing depth and increasing hydrostatic
gen,25 while other microorganisms consume it, usually resulting pressure as well as between increasing pressure and increasing
in nanomolar concentrations of dissolved hydrogen in a steady gas solubility following Henry’s law. Also, an unintended
state.25−41 A competition for hydrogen may then result in the leakage of pure hydrogen gas into a shallow aquifer (up to 140
development of a typical redox sequence (i.e., decreasing redox m depth) will result in a much higher dissolved hydrogen
B DOI: 10.1021/acs.est.7b05467
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

Table 2. Chemical Composition of the Medium Grained Sandy Sedimenta


Nab Kb Cab Mgb Mntotb Fetotb FeIIc FeIIIc Sredd Corge Cinorge
3.68 8.8 23.2 0.797 0.121 3.91 <0.02 0.39 0.49 0.27 4.9
a
Conc. in g/kg. bAfter total digestion of sediment sample in hydrochloric acid, nitric acid, hydrofluoric acid, and perchloric acid. cExtractable after 21
days in 5 M HCl according to the methods of Heron et al.79 dAnalyzed as chromium reducible sulfur (CRS) similarly to the methods of Canfield et
al.80 eAnalyzed with Ströhlein Coulomat Type 702.

Table 3. Composition of the Groundwater and the Inflow Solution of the Column Experimenta
groundwater inflow solution
parameter unit n n
elec. cond. [μS cm−1] 721 ± 9 16 630 ± 38 7
pH [−] 7.2 ± 0.2 16 8.5 ± 0.3 8
O2 [mg L−1] 0.011 ± 0.007 12 n.m.
EHSHE [mV] −122 ± 59 16 n.m.
T [°C] 9.8 ± 0.2 16 20 ± 2
Na+ [mmol L−1] 0.72 ± 0.03 13 0.33 ± 0.08 21
K+ [mmol L−1] 0.04 ± 0.001 13 0.02 ± 0.01 21
Ca2+ [mmol L−1] 3.49 ± 0.06 13 1.2 ± 0.5 21
Mg2+ [mmol L−1] 0.234 ± 0.005 13 0.11 ± 0.03 21
Sitot [mmol L−1] 0.468 ± 0.008 13 0.21 ± 0.03 21
Fetotb [μmol L−1] 32 ± 7 13 0.3 ± 1.2 20
Mntotb [μmol L−1] 3.77 ± 0.06 13 0.5 ± 1.0 20
Cl− [mmol L−1] 0.81 ± 0.03 15 0.315 ± 0.007 22
SO42− [mmol L−1] 0.95 ± 0.04 15 0.33 ± 0.04 22
NO3− [μmol L−1] 6 ± 37 14 0.9 ± 2.0 22
PO43− [μmol L−1] <1.05 14 <1.05 22
F− [μmol L−1] <5.3 14 <5.3 22
Br− [μmol L−1] <1.25 14 <1.25 22
CH3COO− [μmol L−1] <3.33 14 30 ± 38c 25
CH4 [μmol L−1] 3±1 12 <1 20
TIC [mmol L−1] 5.4 ± 0.2 15 1.5 ± 0.2 22
TOCd [mmol L−1] n.m. 0.26 ± 0.09 22
NPOCe [mmol L−1] 0.4 ± 0.1 14 n.m.
a
Mean and standard deviation from n analyses; n.m.: not measured. Note: a table containing the average outflowing concentrations in the same data
structure can be found in the Supporting Information, Supporting Table 1. bDissolved Fe and Mn species were apparently oxidized and precipitated
during the storage of the raw groundwater collected from the field. These precipitates were filtered out when preparing the inflow solution so that
total Fe and Mn concentrations were significantly lower in the inflow solution than in the groundwater. cAcetate has not been added to the inflow
solution or to any other material used for the experiments when preparing the material and the setup. Its appearance in minor quantities is most
probably explained by acetogenic microbial activity taking place in the high-pressure mixing cell. dTOC was calculated as a difference of TC and TIC
measurements. eNPOC: nonpurgeable organic carbon; measurement of TC after acidifying (HCl) and stripping (O2) the sample for 5 min.

concentration than the nanomolar quantities usually present in electron donor limitation, while after a gaseous hydrogen
aquifers (Table 1). Elevated hydrogen concentrations will intrusion into an aquifer the availability of hydrogen as electron
probably result in a surplus in hydrogen supplies for the donor is not limiting any more. The knowledge of establishing
microbial community, which usually strives for hydrogen. the sequence of redox reactions, their reaction kinetics, and
Therefore, the microorganisms may start consuming hydrogen reaction products is necessary for any risk assessment or spatial
rapidly through a series of consequent redox reactions46,63 or planning attempts and for modeling efforts at reservoir or
through simultaneous redox processes64 without the establish- regional scales.7,8 Furthermore, monitoring and operation
ment of typical redox sequences. On the other hand, the strategies for a gas storage site also require knowledge about
groundwater above a gas storage reservoir may have an elevated the fate of the stored gas in shallow aquifers completed by an
salt content and a saltwater or brine intrusion might accompany estimation of the impact the stored gas may have through
a gas leakage into a shallow aquifer. Elevated salinity potentially volatile products of redox reactions (e.g., H2S or CH4).
inhibits the microbiologically catalyzed redox reactions due to a The study presented here shows results of a high-pressure
limited halotolerance of the microorganisms present.65−72 column experiment using natural aquifer sediment and
In order to estimate the potential effects the subsurface groundwater equilibrated with hydrogen partial pressures of
storage of hydrogen could have on groundwater, including the 2−15 bar, representing a scenario of hydrogen gas leakage into
effects following an unintended leakage into a shallow aquifer, a shallow or medium deep aquifer. The results enabled the
an improved understanding of hydrogenotrophic redox development of a simplified reaction model scheme describing
reactions at high dissolved hydrogen concentration is required, simultaneous redox reactions at elevated dissolved hydrogen
but investigations on those systems are missing so far, especially concentrations at flow through conditions, including reaction
regarding reaction kinetics. Available models on the spatial or kinetics and rate coefficients. Furthermore, results of a simple
temporal zonation of redox processes arise in part from testing routine for estimating the halotolerance of a microbial
C DOI: 10.1021/acs.est.7b05467
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

community are shown. The outcome of the study is a first step was kept constant, and after that discontinuity, the groundwater
in an improvement of the understanding of hydrogen oxidizing percolation in the experiment was resumed.
redox reactions in the context of hydrogen gas storage. The A pressure-resistant sampling cell76−78 was used to sample
contributions of this work will also be integrated into the the in- and outflowing solutions as well as solutions from the
Compendium on Concepts and Methods for Quantitatively sampling ports along the flow path without discharging
Deducting the Spatial Demand of Subsurface Energy Storage pressure in the sediment column. After releasing the pressure
prepared by the ANGUS+ research project facilitating future from the sampling cell, the water in it was sampled, treated, and
improvements in planning the sustainable use of the subsurface analyzed for cation, anion, and total as well as inorganic carbon
space.6−8 concentrations. While the above-mentioned parameters could


be determined precisely, taking a sample from a solution
MATERIALS AND METHODS saturated by a p(H2) above the atmospheric pressure resulted in
rapid hydrogen outgassing during pretests. Thus, before
Experimental Setup. Sediment from a pristine shallow samples for hydrogen analyses were extracted from the
Quaternary aquifer (WGS84: N54.0773, E10.0750, Table 2) sampling cell, the water to be sampled was diluted under the
obtained from drilling cores between 4 and 12 mbgs was used column’s in situ water pressure within the sampling cell by
in the experiments. The sediment was homogenized and stored factors between 1:5 and 1:25. The dilution was carried out by
wet in an Ar atmosphere to preserve the redox conditions as pumping well-defined amounts of the hydrogen free initial
well as possible. A pre-experiment conditioned the partially solution into the sampling cell, simultaneously to the percolated
oxidized sediment for approximately two months before the sample. This dilution lowered the hydrogen concentrations in
high-pressure column experiment started using a flow through the sampling cell to below the saturation concentration at 1 bar
batch reactor at atmospherical pressure and enriched by p(H2): this method delivered reasonable hydrogen concen-
hydrogen gas at approximately 1 bar.64 The batch experiments trations and negligible losses in pretests. Pressure resistant
for testing halotolerance contained unconditioned sediment. sensors (Corr Instruments, San Antonio, TX, USA) connected
Groundwater from the same well the sediment cores originated to the sampling cell enabled the measurement of pH and
from was sampled periodically (Table 3) and stored in HDPE electrical conductivity during each sampling event at in situ
canisters before being utilized in the column experiment or in pressure conditions.
gastight Al-impregnated bags (Tesseraux) for batch experi- For the halotolerance test, 115 mL glass serum vials were
ments. Vacuum filtration (cellulose nitrate membrane 0.45 μm, used as microcosms and filled with 50 g of untreated sediment
Sartorius) and dilution using deionized water (1:2.5) of the and 75 g of groundwater. An enrichment of the groundwater
groundwater prevented the formation of precipitates potentially with CaSO4 and NaCl salt adjusted a TDS level between 0.35
clogging the pumps and capillary tubes supplying the column and 350 g L−1 (Table 4). A thorough sparging with pure
experiment. Equilibrating this initial solution of diluted
groundwater to various p(H2) in a flow through mixing cell Table 4. TDS Level and Added Salt Concentration (in mg
directly upstream the column generated in-line the inflow L−1) in the Halotolerance Testing Schedule
solution for the high-pressure column experiment (Table 3).
TDS level CaSO4 NaCl
The basic setup of the high-pressure column experiment
itself is described in detail elsewhere.76−78 In short, the high- 350 0 0
pressure column used (Ø = 4 cm, L = 50 cm, electrochemically 1100 375 375
passivated titanium grade 2, with 4 ports along the flow path, 3500 1575 1575
from Werner Kluge Engineering GmbH, Kiel, Germany) was 11 000 2000 8650
equipped with stainless steel and hastelloy fittings and tubes 35 000 2000 32 650
(Swagelok, OH, USA) allowing an investigation in a gastight 110 000 2000 107 650
system at total pressures of up to 100 bar. One HPLC pump 350 000a 2000 347 650
a
(Knauer GmbH, Berlin, Germany) pumped the initial solution A minor fraction of the added salt did not dissolve.
into a mixing cell where hydrogen (grade 5.0) was added at a
constant pressure monitored in the headspace by a hydrogen gas for several minutes ensured an initial dissolved
precalibrated pressure gauge (Swagelok, OH, USA). A second hydrogen concentration nearly in equilibrium with p(H2) = 1
HPLC pump pumped the generated inflow solution through bar. Furthermore, by sealing the vials using aluminum crimps
the column from the bottom to the top. A variable backpressure containing a Teflon plated butyl rubber septum as fast as
valve at the end of the flow line ensured a total pressure in the possible after the sparging, a nearly pure hydrogen atmosphere
system of up to 10 bar higher than the applied gas partial in the headspace was prepared. The microcosms were stored
pressures in order to prevent degassing in the experimental still and upside down in dark boxes at room temperature (20 ±
apparatus. In the experiments presented here, hydrogen gas at 2 °C), and samples for hydrogen analysis were taken after
p(H2) between 2 and 15 bar was applied during different distinct time intervals from the free water phase after shaking
experimental stages (Supporting Information, Supporting and opening a serum vial. Three separate batches were
Figure 1). A pumping rate of 0.5 mL min−1 created a flow prepared for each sampling event, which took place after 0, 6,
speed of approximately 2 m d−1, assuming an effective porosity 15, 23, and 48 days.
of 0.3 ± 5% based on equivalent experiments,76,77 which results Sample Preparation and Analytics. Gas chromatography
in a total residence time of approximately 6.3 h. The total run (7890B GC with a TCD detector coupled to a 7697A
time of the column experiment was 242 days. A technical issue Headspace Sampler, Agilent Technologies) was used for
with the HPLC pumps interrupted the flow through the hydrogen analytics after a three interval linear calibration
column at day 105 for 40 days between the 2 and 3 bar runs; from 25 μM to 10.5 mM based on defined volumes of
but in this phase, the total pressure in the column experiment hydrogen gas (99.999 vol %) added by Hamilton syringes into
D DOI: 10.1021/acs.est.7b05467
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

closed 20 mL headspace vials containing 5 mL of deionized


water, sealed with a Teflon-coated butyl rubber septum crimp,
stored upside down. Freshly prepared control standards
measured over several months showed a standard deviation
of 7.5% (n = 546). Hydrogen concentrations were measured in
5 mL samples as described for the standards using up to three
parallel and unfiltered samples, typically within an hour after
sampling. All other samples were filtered (0.2 μm, regenerated
cellulose, Sartorius) and filled into vials for cation, anion
(HDPE), and inorganic and organic carbon (glass, sealed with
the same crimps used for hydrogen samples) measurements.
Cation samples were acidified (cc. HNO3 s.p.), and all samples
were stored at 4 °C in the dark until the measurement. An ICP-
AES (Vista AX, Varian) analyzed the total elemental abundance
of Na, K, Ca, Mg, Fe, Mn, and Si. An ion chromatograph
analyzed (Metrohm 881) the anion (formate, acetate, sulfate,
chloride, nitrate, phosphate) concentrations, and a TIC-TOC
analyzer (multi N/C 2000 from Analytik Jena) measured the
total inorganic carbon (IC) and total carbon (TC) as well as
the nonpurgeable organic carbon (NPOC) concentrations, all
at the day of sampling. Ammonium concentrations were
measured photometrically, following DIN 38406E5. The
saturation index for, e.g., calcite was calculated as the ten-
based logarithm of the ratio between the ion activity product of
its dissolved components (IAP) and the corresponding
equilibrium constant for solubility (K) using PHREEQC81,82
and phreeqc.dat completed by a redox decoupled species for
acetate. Data storage and evaluation, including the model fitting
process, was done using MS Excel.


Figure 1. Dissolved concentrations of sulfate, acetate, hydrogen,
calcium, and total inorganic carbon species (IC) as well as the pH
RESULTS AND DISCUSSION value in the column’s in- and outflow solutions versus the experimental
Experimental Results. The column experiment started run time in days. Vertical gray lines show the consecutive p(H2)
using an inflow solution in equilibrium with a p(H2) of 5 bar, changes in the inflowing solution. (A technical note: the “15-bar”
period would only add 3 data points to the “Hydrogen” chart and 2
and after 56 days, the hydrogen partial pressure was increased
points to the “pH” chart, but it would, in exchange, degrade the clarity
to 15 bar for 18 days. Then, after the distribution of of the whole figure. These points were therefore allowed to go off-scale
geochemical parameters in the column had reached steady on those charts instead of removing them from the whole figure.)
state, the partial pressure of hydrogen was decreased back to 5
bar for 17 days followed by experimental phases with even
lower p(H2) of 2 and 3 bar for 16 and 42 days, respectively. These changes in the composition of groundwater were in a
Finally, a third 5-bar period of 22 days completed the good agreement with the assumption of an established
experiment. The results indicated that the flow interruption hydrogenotrophic microbial community apparently oxidizing
of 40 days between the 2-bar and 3-bar periods did not affect hydrogen mainly by reducing sulfate and reducing carbon
the system significantly and, therefore, the time interval of the dioxide through acetogenesis. Moreover, the decrease in
flow interruption was neglected; the results are shown on a calcium concentrations linked to increasing pH values along
continuous time scale (Figure 1). These results show a the flow path suggested the precipitation of calcium carbonate,
continuous change between the in- and outflow regarding which also contributed to the decrease in IC (Figure 2). The
dissolved concentrations of sulfate, acetate, total inorganic dissolved Fetot concentrations stayed below the detection limit
carbon species (IC), calcium, and hydrogen, starting without of 0.02 mg L−1 in all samples while the concentration of
any detectable lag time at the beginning of the experiment. dissolved Mntot exceeded the detection limit of 0.01 mg L−1
Figure 1 demonstrates decreasing concentrations of sulfate (by only in a few samples, suggesting that FeIII and MnIII/IV
160 ± 100 μM; n = 77), calcium (500 ± 400 μM, n = 79), IC reducing processes were of minor importance. However, solely
(800 ± 400 μM, n = 73), and dissolved hydrogen (1700 ± 700 measuring dissolved Fetot concentrations serves only as an
μM, n = 47), while the acetate concentration increased (180 ± indicator for identifying FeIII-reduction processes, because
50 μM, n = 61). Furthermore, a sulfidic smell was clearly precipitating mineral phases (i.e., FeS, FeCO3) may limit the
identifiable in the solution flowing out of the column, and dark solubility of FeII. The same is valid for any MnIII/IV reduction
precipitates, most probably iron and manganese sulfides or processes. As the experiments are ongoing, mineralogical
possibly their carbonates and/or hydroxides, were also evaluation of the solid phase was beyond the subject of this
observed in a transparent tube which discharged the water work and, therefore, the role of FeIII and MnIII/IV reduction is
from the outflow valve. An electron balance can be read from not conclusively evaluated. However, the comparatively small
Figure 4, as the modeled hydrogen concentration reflects values FeIII content of the sediment used (0.55 ± 0.39 g kg−1)
calculated from the stoichiometry of electron consuming sulfate supported the assumptions on a minor importance.
reduction and acetogenesis, which are compared to the The pH usually increased from 8.7 to 9.6 along the flow path
measured hydrogen values. and up to 11.8 during the 15-bar period of the experiment. The
E DOI: 10.1021/acs.est.7b05467
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

the 3-bar period of the experiment. However, as no connection


between the variations in sulfate decrease and the dissolved
hydrogen concentration in the inflow can be seen (Figure 5),
these fluctuations in sulfate reduction rates are possibly caused
by variations in microbial activity not controlled by any of the
monitored parameters, including minor fluctuations in
inflowing sulfate concentrations representing natural changes
in the groundwater-based solution used to percolate the
columns.
The concentration profiles along the flow path gathered
during all six experimental phases (Figure 2) support the
assumption that redox reactions, most probably catalyzed by
microorganisms, were responsible for the observed changes in
dissolved concentrations of hydrogen, sulfate, and acetate.
Moreover, the concentration profiles acknowledged the
independency of the reaction rates from the initial dissolved
hydrogen concentration derived from monitoring the in- and
outflow concentrations (Figure 1). All concentration profiles
show an average decrease in dissolved concentration of sulfate
from 0.33 to 0.22 mM, of IC from 1.45 to 0.70 mM, of calcium
from 1.13 to 0.84 mM, and of hydrogen by 1.2 mM as well as
an average increase in acetate concentration from 0.05 to 0.2
mM slowing down toward the end of the flow path. The results
imply that all reactions started immediately and concurrently
along the whole flow path. The simultaneous reaction progress
established throughout the total length of the column indicated
that a classical redox sequence, i.e., subsequent and separate
sulfate reducing and acetogenic zones, was not established in
the column experiment exhibiting hydrogen surplus at any time.
Figure 2. Dissolved hydrogen, sulfate, total inorganic carbon species, Furthermore, the almost linear decrease in hydrogen, sulfate,
acetate, and calcium concentrations, pH, and electrical conductivity as IC, and calcium concentration suggests that the reactions
well as the saturation index81,82 of calcite for the solutions percolating responsible for the concentration changes proceeded more or
the high-pressure column plotted versus the residence time in the less evenly along the total length of the flow path. Only the
experiment. Different profile lines represent sample series taken during acetate formation rate decreased toward the end of the column.
consecutive experimental phases using diluted groundwater equili- The concentration profiles provided an estimation of an
brated with different hydrogen partial pressures as inflowing solution
average net hydrogen consumption rate of 0.25 mM h−1, which
(e.g., “5 bars (2nd profile)”) represents the second experimental phase
with 5 bar hydrogen partial pressure; see Figure 1. H2 in 15 bar was nearly balanced by the 0.2 mM h−1 that was calculated on
experiment not depicted. the basis of the stoichiometry of hydrogenotrophic sulfate
reduction and acetogenesis. Although their rates could not be
quantified, FeIII and MnIII/IV reduction most probably also
concentrations of other groundwater constituents (i.e., sodium, contributed to the hydrogen consumption and they would
potassium, magnesium, chloride) showed no mentionable therefore explain a part of the hydrogen surplus. The rest of the
differences between the in- and outflows of the column gap in this redox balance was probably caused by a loss of
experiment, apart from an unsteady Sitot increase (50 ± 80 μM, hydrogen during the sampling procedure or by a diffusive loss
n = 62). The increase of silica concentrations can indicate that through pumping heads, tubes, connectors, etc. of the system
weathering of silicate aquifer minerals took place, possibly apparatus, reflecting the serious challenge to work with the
enabled by the increasing pH. An overall consequence of the highly fugitive hydrogen gas dissolved at partial pressures much
observed hydrogeochemical processes was a decrease in the higher than the ambient pressure. Hydrogen consuming
electrical conductivity of the groundwater by 150 μS cm−1. processes other than the discussed ones cannot be totally
The composition of the inflow solution varied slightly due to excluded, but the results do not suggest any other relevant
a repeated renewal of the stock in freshly produced ground- reactions.
water that was used for its preparation, reflecting the Methane (detection limit: 1 μM) or formate (detection limit:
background fluctuation of a natural aquifer. The variable H2 4 μM) production were not detected in any of the samples,
inflow concentration caused by different applied p(H2) shows although they were expected.83 The increasing pH value from
no correlation with these fluctuations. Furthermore, the extent 8.7 to 9.6 along the flow path of the column experiment might
of concentration changes in dissolved species between the in- have inhibited methane formation because pH levels above 7−9
and outflow generally shows no correlation to the hydrogen can prevent methanogenic activity84,85 although studies at town
inflow concentration, and variations in these changes did not gas storage sites and hydrogen-limited aquifers have shown
show a clear trend either. An exception is the loss in sulfate methane formation at conditions comparable to the column
concentration, where a temporarily decreasing tendency was experiment.14,17−19 Usually, methane production is known to
observed between day 56 (ΔSO42− = 80 μM) and day 78 take place at rates of up to a third of the acetate production
(ΔSO42− = 22 μM), followed by a reestablishment to initial rates32 and even the presence of acetate itself in the column
values and then a slight increasing tendency at the beginning of experiment made methane production expectable.86 Therefore,
F DOI: 10.1021/acs.est.7b05467
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

the high pH value characterizing the experiment probably in consumption rates in the different experimental series.
combination with a lack of specialized microbes could be the Hydrogen consumption being inhibited by increasing the
reason for suppressed methane formation. However, the salinity acknowledged that microbial processes were responsible
continuation of the lab experiment at neutral pH values for hydrogen oxidation. This further suggests that TDS may
allowed methanogenesis to start.87 Interestingly, earlier control or limit hydrogen oxidation, if microbial activity takes
predictions assumed a groundwater acidification as a con- place. Such a TDS limit will probably depend on the site-
sequence of a surplus in molecular hydrogen,21 which is clearly specific microbial community of an aquifer.
denied by the presented experimental results. The microbially mediated redox reactions starting sponta-
Overall, the column experiment represents a flow through neously in the preconditioned sediment used in the column
system characterized by the surplus of dissolved hydrogen and experiment and after a short lag time in the halotolerance test
dominated by spontaneously starting hydrogen oxidizing suggest that the unspecified natural microbial communities in
reactions, mainly parallel sulfate reduction and acetogenesis the aquifer material were ready to use high concentrations of
catalyzed by microorganisms. Dissolved hydrogen was removed hydrogen as electron donor as well as sulfate and inorganic
almost completely during the 2-bar period of the experiment carbon as electron acceptors. Furthermore, the results show
only, indicated by a nonlinear decrease in hydrogen that at p(H2) higher than 2 bar the increasing dissolved
concentration along the flow path (Figure 2) and small hydrogen concentration had no effect on the type nor on the
concentrations of hydrogen in the outflow (between 27 and velocity of the established redox reactions. The results of the
158 μM) (Figure 1). However, the temporary depletion of halotolerance test combined with the column experiment
dissolved hydrogen did not stop acetate production, pH supported the idea of a fast establishment of microbially
increase, or calcium precipitation. The lack of dependency of catalyzed redox reactions in case of a hydrogen gas leakage into
the observed reactions on the initial p(H2) or actually dissolved shallow or medium deep aquifers, even in a scenario including
H2 concentration indicated that the system reached quasi salt water intrusion accompanying the gas leakage. Storage of
steady-state conditions. Thus, the established microbial hydrogen gas in porous saline aquifers will probably also initiate
community was apparently not sensitive to the pressure these redox reactions, if the salt content in the aquifer does not
variations or changing dissolved H2 concentration applied in exceed some tens of grams per liter.
the experiment at least as long as a surplus of dissolved Kinetics. Two process models using different approaches
hydrogen was provided. were consecutively developed for describing the observed
The results of complementary batch experiments applying a reactions. First, dual substrate Monod kinetics was initially
range of different total dissolved solids (TDS) concentrations parametrized on the basis of literature values and it was then
showed that high concentration of TDS inhibited hydrogen- fitted to the experimental results by varying the maximum
consuming redox reactions (Figure 3). The microcosms reaction rates and the Monod half-saturation constants (Table
5, Figure 4) for describing microbial metabolic processes
potentially limited by substrate availability. However, Monod
kinetics appears as either zeroth or first order kinetics in the
case that steady-state microbial activity can be assumed and if
reactant concentrations are significantly higher or lower than
the Monod half-saturation constants, respectively.45,88,89 Thus,
a simplified combination of (pseudo) zeroth and (pseudo) first
order rate equations and stoichiometric balance equations
represented the other kinetic process modeling approach
discussed here.
The first modeling approach based on dual substrate Monod
kinetics81 can describe the experimental data as illustrated in
Figure 4 using eqs 1−3 with the parameter values fitted to the
Figure 3. Development of dissolved hydrogen concentrations experimental results (Table 5). Thereby, consumption of
normalized to the initial total moles of hydrogen at different total dissolved hydrogen and IC were defined on the basis of
dissolved solids (TDS, mg L−1) concentrations. The black line stoichiometric balance equations using eqs 4 and 5,
represents a blank series with quartz sand and deionized water; all respectively. Thus, unspecific hydrogen loss in the experimental
standard deviation bars were calculated from triplicate samples. Note
setup was not overinterpreted and carbonate precipitation,
that the blank series identified no major artifacts or experimental side
effects (e.g., diffusive losses, sorption, etc.). which could not be quantified due to the lack of mineral
analyses, was regarded.

containing groundwater with TDS higher than 35 g L−1, d[Ca 2 +]


− = kPrec
which corresponds to the average salinity of seawater, showed dt (1)
constant hydrogen concentrations during a time span of at least
48 days. These concentrations were equivalent to the blank d[SO4 2 −] [H 2] [SO4 2 −]
experiments containing quartz sand and hydrogen-saturated − = vmaxSR × ×
dt MK H2−SR + [H 2] MK SO4 2 − + [SO4 2 −]
deionized water. In all microcosms containing TDS of 11 g L−1
(2)
or less, a significant and similar decrease in hydrogen
concentration was observed after a lag time of less than 6
days. The relative concentration of dissolved hydrogen reached d[Ac−] [H 2] [IC]
− = vmaxAcG × ×
values between 60% and 30% after 48 days in these batch series. dt MK H2−AcG + [H 2] MK IC + [IC]
Thereby, all these microcosms exhibited comparable hydrogen (3)

G DOI: 10.1021/acs.est.7b05467
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

Table 5. Literature and Experimental Parameters for the Calculation of the Monod Kinetics Regarding Sulfate Reduction (SR)
and Acetogenesis (AcG), Such as the Maximal Reaction Rate (vmax) and the Monod Half-Saturation Constants (MK) for the
Corresponding Substratesa
vmax SR (h−1) MK SO42− (mM) MK H2−SR (mM) vmax AcG (h−1) MK IC (mM) MK H2−AcG (mM)
Literature Values
Boudreau and Westrich44 1.6
Dale, Regnier, and Van Cappellen45 0.002988 1 0.00001 0.002052 0.0001
Lovley and Goodwin46 0.0014
Watson, Oswald, Mayer, Wu, and Banwart47 0.000324 0.16 0.001
Xu, Jaffé, and Mauzerall48 0.23 0.0029
selected starting value before fitting process 0.002988 1 0.00001 0.002052 3 0.0001

Fitted to Experimental Result


2 bar model 0.05 1 0.00001 0.55 17 0.0001
3 bar model 0.12 1 0.00001 0.6 17 0.0001
5 bar (1st model) 0.072 1 0.00001 0.5 17 0.0001
5 bar (2nd model) 0.075 1 0.00001 0.55 17 0.0001
5 bar (3rd model) 0.12 1 0.00001 0.6 17 0.0001
15 bar model 0.052 1 0.00001 0.27 17 0.0001
averaged results 0.07 ± 0.04 1 0.00001 0.5 ± 0.2 17 0.0001
a
Note that all MK values for hydrogen are 4 to 5 orders of magnitude lower than the dissolved hydrogen concentrations in the high-pressure
experimental columns (Figure 2), representing concentrations expected to take place in an aquifer after a gaseous hydrogen intrusion took place.

Figure 4. Results of the simplified models using first order kinetics for acetate production and zeroth order kinetics for sulfate reduction and Ca-
carbonate precipitation (colored lines) and results of the model based on dual substrate Monod kinetics (black dotted lines lying perfectly on the
colored lines) and measured dissolved concentrations of hydrogen, sulfate, IC, calcium, and acetate (symbols) along the flow path profile in the high-
pressure column experiment during different experimental phases. The residence time was calculated using mean flow and porosity; dispersion was
neglected.

d[H 2] d[Ac−] d[SO4 2 −] Following the second-simplifying-modeling approach, sulfate


− =4 +4 reduction was calculated by zeroth order kinetics (eq 6) and
dt dt dt (4)
acetogenesis was described by pseudo first order kinetics in
dependence of IC (eq 7). Hydrogen and IC decrease as well as
d[IC] d[Ac−] d[Ca 2 +] carbonate precipitation were calculated in the same way as it
− =2 +
dt dt dt (5) had been done in the Monod kinetic modeling approach (eqs 4,
H DOI: 10.1021/acs.est.7b05467
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

5, and 1, respectively). The kinetic rate constants for sulfate several orders of magnitude higher than the corresponding
reduction (kSR), acetogenesis (kAcG), and carbonate precip- Monod half-saturation constants. Moreover, no literature values
itation (kPrec) were fitted to concentration profiles of Ca2+, are known for the half-saturation coefficient “MK IC”. Overall,
SO42−, acetate ([Ac−]), IC, and dissolved hydrogen. This for further applications, we suggest the more robust assumption
simplified model was equally able to represent the concen- of (pseudo) first and zeroth order kinetics, as these were
tration profiles (Figure 4) using rate coefficients for sulfate directly supported by the experimental results. For including
reduction and acetogenesis between 9 and 26 μM h−1 (average further or different pathways of the reduction of inorganic
18 ± 5) and 0.015 and 0.033 h−1 (average 0.030 ± 0.006), carbon species, eqs 4 and 5 can be extended by terms
respectively (Table 6). The fitted calcite precipitation rate representing reactions such as methane or formate production.
coefficients varied between 39 and 88 μM h−1 (average 57 ± Figure 5 shows the (pseudo) first and zeroth order rate
17). coefficients for acetogenesis, sulfate reduction, and carbonate
precipitation, respectively, gathered by fitting the model and
d[SO4 2 −]
− = k SR plotted against the initial dissolved hydrogen concentration.
dt (6) For the rate coefficients, the model input data were the six
concentration profiles along the flow path as well as all
d[Ac−] concentration differences between the inflow and outflow
− = kAcG × [IC]
dt (7) solution. The averaged values considering the in- and outflow
concentrations from each experimental period are mostly in
Table 6. (Pseudo) Zeroth Order Rate Coefficients for good agreement with the results provided by the concentration
Hydrogenotrophic Sulfate Reduction (kSR) and calcium profiles. Overall, the rate coefficients may vary by a factor of 2
carbonate precipitation (kPrec) and Pseudo First Order Rate to 3, but they indicated no dependency on the initial hydrogen
Coefficient with Respect to IC for Acetogenesis (kAcG) concentration, suggesting a quasi-steady state of the microbial
kSR (μM·h−1) kAcG (h−1) kPrec (μM·h−1)
community. Moreover, no recognizable trend in the microbial
activity versus the experimental run time (Figure 1) suggests no
average 18 ± 5 0.030 ± 0.006 57 ± 17
overall growth in the active biomass. The mean specific sulfate
reduction rate of 42 nM·cmsediment−3·h−1 or the mean
2 bar model 9 0.030 55
3 bar model 24 0.033 72
acetogenesis of 85 nM·cmsediment−3·h−1, which were calculated
5 bar (1st model) 17 0.028 40
from the column experiment, are smaller than rates described
5 bar (2nd model) 17 0.030 47
for other reactive aqueous environments exhibiting high
5 bar (3rd model) 26 0.033 88 dissolved hydrogen concentrations. Experimental studies58,90,91
15 bar model 13 0.015 39 on bioreactors yielded sulfate reduction rates between 3060
and 79 500 nM·cmsediment−3·h−1. However, a study90 on
acetogenesis in a Cretaceous shale and sandstone environment
The model using the combination of (pseudo) first or zeroth showed rates reaching from 110 to 160 nM·cmsediment−3·h−1
order reaction kinetics exactly reproduced the model curves parallel to sulfate reduction. While the acetogenesis rate was at
calculated using Monod kinetics and, therefore, both modeling least in a similar range, the sulfate reduction was 2 or 3 orders
approaches could be equally suitable. However, the Monod of magnitude smaller in the column experiment. Several reasons
kinetics model is strongly under-parametrized due to the might be responsible for the comparatively low metabolic
comparatively small changes in the total dissolved inorganic activity, e.g., a lack of nutrients (Table 3), composition of the
carbon species and sulfate concentrations as well as due to the established microbial community, or the comparatively high pH
constant availability of hydrogen in concentrations that were value.

Figure 5. Rate coefficients describing sulfate reduction, acetate formation, and calcium carbonate precipitation in the column experiment versus the
initial dissolved concentration of hydrogen. Vertical and horizontal error bars represent standard deviations in the measured hydrogen
concentrations and in reaction rate constants fitted for every in- and outflow concentration (“_in−out”), respectively. The larger symbols (“_port”)
represent the six fits based on concentrations over the flow path (Table 5, Figure 4).

I DOI: 10.1021/acs.est.7b05467
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

The Monod model as well as the simplified zeroth and first groundwater and the stored gas. Therefore, targets of a
order model supported the assumption that the simultaneously monitoring network covering a hydrogen gas storage site in
proceeding and microbially catalyzed hydrogenotrophic sulfate order to detect a gas leakage should be focused on the effects
reduction and acetogenesis were the main redox processes reactions can have on groundwater (e.g., appearance of acetate,
established in the column experiment, accompanied by depletion of sulfate, decreasing electric conductivity) instead of
carbonate precipitation induced by the increasing pH value. the stored gas initially dissolved in a shallow aquifer.
Nevertheless, sulfate reduction by acetate oxidation is also a
well-known microbial metabolic process86 and a model
considering this process instead of direct hydrogenotrophic

*
ASSOCIATED CONTENT
S Supporting Information
sulfate reduction could also be fitted to the results of the The Supporting Information is available free of charge on the
column experiment. However, such a model would be under- ACS Publications website at DOI: 10.1021/acs.est.7b05467.
parametrized using the available data set, because acetotrophic
sulfate reduction would consume acetate and produce carbon Table of experimental parameters; average hydrogen
dioxide, which is then reduced again to acetate by oxidizing concentrations in different experimental runs (PDF)


hydrogen. Consequently, the acetate oxidation rate would be
twice as high as the sulfate reduction rate for stoichiometric AUTHOR INFORMATION
reasons and the rate of acetogenesis would increase by the same Corresponding Authors
value, resulting in identical model results. The experimental
*E-mail: mb@gpi.uni-kiel.de (M.B.).
results support a parallel and not a sequential occurrence of
*E-mail: frank.dethlefsen@ifg.uni-kiel.de (F.D.).
acetogenesis and sulfate reduction, as both processes started
immediately at the beginning of the flow path, thereby ORCID
providing no evidence for sulfate reducing processes other Marton Berta: 0000-0001-8467-2792
than hydrogenotrophic sulfate reduction. However, an Notes
appearance of acetotrophic sulfate reduction cannot be totally The authors declare no competing financial interest.


ruled out.
Implications. The main purpose of this study is to support ACKNOWLEDGMENTS
the planning, operation, and monitoring of hydrogen storage
facilities as well as to contribute to planning the sustainable use We would like to thank the funding provided by the German
of the subsurface space. An important contribution to these Federal Ministry of Education and Research (BMBF) for the
workflows is the investigation of a hydrogen-rich groundwater ANGUS+ project (03EK3022), as well as the support of the
environment in which simultaneously proceeding redox Project Management Jülich (PTJ). Furthermore, we would like
reactions were identified, instead of a redox sequence consisting to thank the five anonymous reviewers for suggestions to
improve this paper.


of separate metabolic zones. This poses a sharp contrast
compared to the hydrogen-limited groundwaters in pristine or
contaminated aquifers known from the literature so far. The REFERENCES
presented findings can be relevant for large scale modeling (1) Sterner, M.; Stadler, I. Energiespeicher − Bedarf, Technologien,
efforts, design of monitoring networks at hydrogen gas storage Integration; Springer-Verlag: Berlin Heidelberg, 2014; p 758.
sites, or laboratory investigations serving a site specific risk (2) Pfeiffer, W. T.; Al Hagrey, S. A.; Köhn, D.; Rabbel, W.; Bauer, S.
Porous media hydrogen storage at a synthetic, heterogeneous field site
assessment. Using pressure-independent elementary kinetics
- Numerical simulation of storage operation and geophysical
instead of sophisticated Monod rate equations in large scale monitoring. Environ. Earth Sci. 2016, 75, 1177.
transport reaction models will probably save computational (3) Taylor, J. B.; Alderson, J. E. A.; Kalyanam, K. M.; Lyle, A. B.;
capacity and may enable easier scenario modeling. Zeroth or Phillips, L. A. Technical and economic-assessment of methods for the
simple first order kinetics might overestimate the metabolic rate storage of large quantities of hydrogen. Int. J. Hydrogen Energy 1986,
in the case of exhausting reaction partners, but this effect 11 (1), 5−22.
should be small at conditions with high initial reactant (4) Evans, D. J. An appraisal of underground gas storage technologies
concentrations. Furthermore, the independence of the observed and incidents, for the development of risk assessment methodology; RR
redox reactions from the dissolved hydrogen concentration 605; Crown: Norwich, 2008.
implies in a converse argument that laboratory investigations on (5) Gao, D.; Jiang, D. F.; Liu, P.; Li, Z.; Hu, S. G.; Xu, H. An
integrated energy storage system based on hydrogen storage: Process
hydrogen leakage scenarios can be performed at system configuration and case studies with wind power. Energy 2014, 66,
pressures lower than in situ conditions, which will drastically 332−341.
reduce laboratory efforts. Thereby, the presented halotolerance (6) Bauer, S.; Beyer, C.; Dethlefsen, F.; Dietrich, P.; Duttmann, R.;
test provides an easy procedure to estimate the effect of a salt Ebert, M.; Feeser, V.; Gorke, U.; Kober, R.; Kolditz, O.; Rabbel, W.;
water intrusion accompanying a hydrogen leakage into shallow Schanz, T.; Schafer, D.; Würdemann, H.; Dahmke, A. Impacts of the
or medium deep aquifers, if the natural material from a use of the geological subsurface for energy storage: an investigation
particular site is available. The results of the halotolerance test concept. Environ. Earth Sci. 2013, 70 (8), 3935−3943.
should not be transferred directly to any other groundwater (7) Bauer, S.; Pfeiffer, T.; Boockmeyer, A.; Dahmke, A.; Beyer, C.
environment, because deep saline aquifers might have a well- Quantifying induced effects of subsurface renewable energy storage.
Energy Procedia 2015, 76, 633−641.
adapted and highly halotolerant microbial community possibly
(8) Kabuth, A.; Dahmke, A.; Beyer, C.; Bilke, L.; Dethlefsen, F.;
enabling the consumption of newly introduced hydrogen even Dietrich, P.; Duttmann, R.; Ebert, M.; Feeser, V.; Gorke, U. J.; Kober,
if the salinity of the groundwater in such an aquifer is higher R.; Rabbel, W.; Schanz, T.; Schäfer, D.; Würdemann, H.; Bauer, S.
than 35 g L−1. However, in less saline aquifers, a rapid Energy storage in the geological subsurface: dimensioning, risk analysis
hydrogenotrophic sulfate reduction as well as a CO2 reduction and spatial planning: the ANGUS plus project. Environ. Earth Sci.
has to be expected, and the reaction products will modify the 2017, 76, 23.

J DOI: 10.1021/acs.est.7b05467
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

(9) Dethlefsen, F.; Beyer, C.; Feeser, V.; Köber, R. Parameterizability discrete zones of high-Iron ground-water. Groundwater 1992, 30 (1),
of processes in subsurface energy and mass storage. Environ. Earth Sci. 29−36.
2016, 75 (10), 1−25. (28) Jakobsen, R.; Albrechtsen, H.-J.; Rasmussen, M.; Bay, H.; Bjerg,
(10) United Nations. Report of the Conference of the Parties on its P. L.; Christensen, T. H. H2 concentrations in a landfill leachate plume
twenty-first session, held in Paris from 30 November to 13 December 2015; (Grindsted, Denmark): in situ energetics of terminal electron acceptor
FCCC/CP/2015/10; United Nations Framework Convention on processes. Environ. Sci. Technol. 1998, 32 (14), 2142−2148.
Climate Change: Paris−Le Bourget, France, 2016; p 36. (29) Kotelnikova, S. Microbial production and oxidation of methane
(11) International Renewable Energy Agency. Perspectives for the in deep subsurface. Earth-Sci. Rev. 2002, 58 (3−4), 367−395.
energy transition: Investment needs for a low-carbon energy system; (30) Chapelle, F. H.; O’Neill, K.; Bradley, P. M.; Methe, B. A.; Ciufo,
International Renewable Energy Agency (IRENA): Abu Dhabi, 2017; S. A.; Knobel, L. L.; Lovley, D. R. A hydrogen-based subsurface
p 204. microbial community dominated by methanogens. Nature 2002, 415
(12) UBA. Energieziel 2050:100% Strom aus erneuerbaren Quellen; (6869), 312−315.
Umweltbundesamt: Dessau-Roßlau, Germany, 2010; p 196. (31) Conrad, R.; Goodwin, S.; Zeikus, J. G. Hydrogen metabolism in
(13) Audigane, P.; Ebrahimiyekta, A.; Pichavant, M. Evaluation of a mildly acidic lake sediment (Knaack Lake). FEMS Microbiol. Lett.
hydrogen migration and geochemical reactivity into underground. In 1987, 45 (4), 243−249.
6th International Conference on Fundamentals & Development of Fuel (32) Conrad, R. Contribution of hydrogen to methane production
Cells, Toulouse, France, 2015. and control of hydrogen concentrations in methanogenic soils and
(14) Panfilov, M. Underground and pipeline hydrogen storage. In sediments. FEMS Microbiol. Ecol. 1999, 28 (3), 193−202.
Compendium of Hydrogen Energy; Gupta, R., Basile, A., Veziroğlu, T. (33) Lovley, D. R.; Chapelle, F. H.; Woodward, J. C. Use of dissolved
N., Eds.; Woodhead Publishing: Oxford, 2016; pp 91−115. H2 concentrations to determine distribution of microbially catalyzed
(15) Bünger, U.; Michalski, J.; Crotogino, F.; Kruck, O. 7-Large-scale redox reactions in anoxic groundwater. Environ. Sci. Technol. 1994, 28
underground storage of hydrogen for the grid integration of renewable (7), 1205−1210.
energy and other applications. In Compendium of Hydrogen Energy; (34) Chapelle, F. H.; Haack, S. K.; Adriaens, P.; Henry, M. A.;
Woodhead Publishing: Oxford, 2016; pp 133−163. Bradley, P. M. Comparison of eH and H2 measurements for delineating
(16) Gniese, C.; Bombach, P.; Rakoczy, J.; Hoth, N.; Schlömann, M.; redox processes in a contaminated aquifer. Environ. Sci. Technol. 1996,
Richnow, H.-H.; Krüger, M. Relevance of deep-subsurface micro- 30 (12), 3565−3569.
biology for underground gas storage and geothermal energy (35) Heimann, A. C.; Blodau, C.; Postma, D.; Larsen, F.; Viet, P. H.;
production. In Geobiotechnology II: Energy resources, subsurface Nhan, P. Q.; Jessen, S.; Duc, M. T.; Hue, N. T. M.; Jakobsen, R.
technologies, organic pollutants and mining legal principles; Schippers, Hydrogen thresholds and steady-state concentrations associated with
A., Glombitza, F., Sand, W., Eds.; Springer Berlin Heidelberg: Berlin, microbial arsenate respiration. Environ. Sci. Technol. 2007, 41 (7),
Heidelberg, 2014; pp 95−121. 2311−2317.
(17) Amigáň, P.; Greksák, M.; Kozánková, J.; Buzek, F.; Onderka, V.; (36) Jørgensen, B. B. Bacteria and Marine Biogeochemistry. In
Wolf, I. Methanogenic bacteria as a key factor involved in changes of Marine Geochemistry; Schulz, H. D., Zabel, M., Eds.; Springer Berlin
town gas stored in an underground reservoir. FEMS Microbiol. Lett. Heidelberg: Berlin, Heidelberg, 2000; pp 173−207.
1990, 73 (3), 221−224. (37) Smith, N. J. P.; Shepherd, T. J.; Styles, M. T.; Williams, G. M.
(18) Panfilov, M. Underground storage of hydrogen: in situ self- Hydrogen exploration: a review of global hydrogen accumulations and
organisation and methane generation. Transp. Porous Media 2010, 85 implications for prospective areas in NW Europe. In Petroleum geology:
(3), 841−865. north-west Europe and global perspectives: proceedings of the 6th
(19) Buzek, F.; Onderka, V.; Vancura, P.; Wolf, I. Carbon-isotope Petroleum Geology Conference; Dore, A. G., Vining, B. A., Eds.;
study of methane production in a town gas-storage reservoir. Fuel Geological Society of London: London, 2005; pp 349−358.
1994, 73 (5), 747−752. (38) Telling, J.; Boyd, E. S.; Bone, N.; Jones, E. L.; Tranter, M.;
(20) Morozova, D.; Zettlitzer, M.; Let, D.; Würdemann, H. MacFarlane, J. W.; Martin, P. G.; Wadham, J. L.; Lamarche-Gagnon,
Monitoring of the microbial community composition in deep G.; Skidmore, M. L.; Hamilton, T. L.; Hill, E.; Jackson, M.; Hodgson,
subsurface saline aquifers during CO2 storage in Ketzin, Germany. D. A. Rock comminution as a source of hydrogen for subglacial
Energy Procedia 2011, 4, 4362−4370. ecosystems. Nat. Geosci. 2015, 8 (11), 851−855.
(21) Reitenbach, V.; Ganzer, L.; Albrecht, D.; Hagemann, B. (39) Lin, L.-H.; Slater, G. F.; Sherwood Lollar, B.; Lacrampe-
Influence of added hydrogen on underground gas storage: a review Couloume, G.; Onstott, T. C. The yield and isotopic composition of
of key issues. Environ. Earth Sci. 2015, 73 (11), 6927−6937. radiolytic H2, a potential energy source for the deep subsurface
(22) Evans, D. J. A review of underground fuel storage events and biosphere. Geochim. Cosmochim. Acta 2005, 69 (4), 893−903.
putting risk into perspective with other areas of the energy supply (40) Cordruwisch, R.; Seitz, H. J.; Conrad, R. The capacity of
chain. In Underground gas storage: worldwide experiences and future hydrogenotrophic anaerobic-bacteria to compete for traces of
development in the UK and Europe; Evans, D. J., Chadwick, R. A., Eds.; hydrogen depends on the redox potential of the terminal electron-
The Geological Society: London, 2009; Vol. 313, pp 173−216. acceptor. Arch. Microbiol. 1988, 149 (4), 350−357.
(23) Truche, L.; Jodin-Caumon, M. C.; Lerouge, C.; Berger, G.; (41) Hoehler, T. M.; Alperin, M. J.; Albert, D. B.; Martens, C. S.
Mosser-Ruck, R.; Giffaut, E.; Michau, N. Sulphide mineral reactions in Field and laboratory studies of methane oxidation in an anoxic marine
clay-rich rock induced by high hydrogen pressure. Application to sediment: Evidence for a methanogen-sulfate reducer consortium.
disturbed or natural settings up to 250 degrees C and 30 bar. Chem. Global Biogeochemical Cycles 1994, 8 (4), 451−463.
Geol. 2013, 351, 217−228. (42) Gujer, W.; Zehnder, A. J. B. Conversion processes in anaerobic
(24) German Drinking Water Act. Trinkwasserverordnung, Verordnung digestion. Water Sci. Technol. 1983, 15 (8−9), 127.
über die Qualität von Wasser für den menschlichen Gebrauch In BGBl. I S. (43) Rotiroti, M.; Jakobsen, R.; Fumagalli, L.; Bonomi, T.
2977 and 3154, 2001. Considering a threshold energy in reactive transport modeling of
(25) Hoehler, T. M.; Alperin, M. J.; Albert, D. B.; Martens, C. S. microbially mediated redox reactions in an arsenic-affected aquifer.
Thermodynamic control on hydrogen concentrations in anoxic Water 2018, 10 (1), 90.
sediments. Geochim. Cosmochim. Acta 1998, 62 (10), 1745−1756. (44) Boudreau, B. P.; Westrich, J. T. The dependence of bacterial
(26) Zinder, S. H. Physiological ecology of methanogens. In sulfate reduction on sulfate concentration in marine sediments.
Methanogenesis: Ecology, Physiology, Biochemistry & Genetics; Ferry, J. Geochim. Cosmochim. Acta 1984, 48 (12), 2503−2516.
G., Ed.; Springer US: Boston, MA, 1993; pp 128−206. (45) Dale, A. W.; Regnier, P.; Van Cappellen, P. Bioenergetic
(27) Chapelle, F. H.; Lovley, D. R. Competitive-exclusion of sulfate controls on anaerobic oxidation of methane (AOM) in coastal marine
reduction by Fe(III)-reducing bacteria - a mechanism for producing sediments: a theoretical analysis. Am. J. Sci. 2006, 306 (4), 246−294.

K DOI: 10.1021/acs.est.7b05467
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

(46) Lovley, D. R.; Goodwin, S. Hydrogen concentrations as an functioning of salt lake ecosystems. Hydrobiologia 2001, 466 (1−3),
indicator of the predominant terminal electron-accepting reactions in 61−72.
aquatic sediments. Geochim. Cosmochim. Acta 1988, 52 (12), 2993− (66) Omar, S. A.; Abdelsater, M. A.; Khallil, A. M.; Abdalla, M. H.
3003. Growth and enzyme-activities of Fungi and Bacteria in soil salinized
(47) Watson, I. A.; Oswald, S. E.; Mayer, K. U.; Wu, Y.; Banwart, S. with sodium-chloride. Folia Microbiol. 1994, 39 (1), 23−28.
A. Modeling kinetic processes controlling hydrogen and acetate (67) Wichern, J.; Wichern, F.; Joergensen, R. G. Impact of salinity on
concentrations in an aquifer-derived microcosm. Environ. Sci. Technol. soil microbial communities and the decomposition of maize in acidic
2003, 37 (17), 3910−3919. soils. Geoderma 2006, 137 (1−2), 100−108.
(48) Xu, S.; Jaffé, P. R.; Mauzerall, D. L. A process-based model for (68) Dincer, A. R.; Kargi, F. Salt inhibition kinetics in nitrification of
methane emission from flooded rice paddy systems. Ecol. Modell. 2007, synthetic saline wastewater. Enzyme Microb. Technol. 2001, 28 (7−8),
205 (3−4), 475−491. 661−665.
(49) Ebert, M. Elementares Eisen in permeablen reaktiven Barrieren zu (69) Uygur, A.; Kargi, F. Salt inhibition on biological nutrient
in-situ Grundwassersanierung − Kenntnisstand nach zehn Jahren removal from saline wastewater in a sequencing batch reactor. Enzyme
Technologieentwicklung. Habilitation Thesis, Kiel University, Kiel, Microb. Technol. 2004, 34 (3−4), 313−318.
Germany, 2004. (70) Yan, N.; Marschner, P.; Cao, W.; Zuo, C.; Qin, W. Influence of
(50) Reardon, E. J. Zerovalent irons: styles of corrosion and inorganic salinity and water content on soil microorganisms. International Soil
control on hydrogen pressure buildup. Environ. Sci. Technol. 2005, 39
and Water Conservation Research 2015, 3 (4), 316−323.
(18), 7311−7317.
(71) Windey, K.; De Bo, I.; Verstraete, W. Oxygen-limited
(51) Schütz, M. K.; Libert, M.; Schlegel, M. L.; Lartigue, J.-E.;
autotrophic nitrification−denitrification (OLAND) in a rotating
Bildstein, O. Dissimilatory iron reduction in the presence of hydrogen:
biological contactor treating high-salinity wastewater. Water Res.
a case study of microbial activity and nuclear waste disposal. Procedia
Earth Planet. Sci. 2013, 7, 409−412. 2005, 39 (18), 4512−4520.
(52) Sandefur, C. A.; Koenigsberg, S. S. The use of hydrogen release (72) Oren, A. Microbial life at high salt concentrations: phylogenetic
compound for the accelerated bioremediation of anaerobically and metabolic diversity. Saline Syst. 2008, 4 (1), 2.
degradable contaminants: the advent of time-release electron donors. (73) Thauer, R. K.; Jungermann, K.; Decker, K. Energy conservation
Remediation Journal 1999, 10 (1), 31−53. in chemotrophic anaerobic bacteria. Bacteriol Rev. 1977, 41 (1), 100−
(53) Ergas, S. J.; Shumway, L.; Fitch, M. W.; Neemann, J. J. 80.
Membrane process for biological treatment of contaminated gas (74) Conrad, R.; Wetter, B. Influence of temperature on energetics of
streams. Biotechnol. Bioeng. 1999, 63 (4), 431−441. hydrogen metabolism in homoacetogenic, methanogenic, and other
(54) Ergas, S. J.; Reuss, A. F. Hydrogenotrophic denitrification of anaerobic-bacteria. Arch. Microbiol. 1990, 155 (1), 94−98.
drinking water using a hollow fibre membrane bioreactor. J. Water (75) Stevens, T. O.; Mckinley, J. P. Lithoautotrophic microbial
Supply Res. Technol. 2001, 50 (3), 161−171. ecosystems in deep basalt aquifers. Science 1995, 270 (5235), 450−
(55) Smith, R. L.; Ceazan, M. L.; Brooks, M. H. Autotrophic, 454.
hydrogen-oxidizing, denitrifying bacteria in groundwater, potential (76) Haase, C.; Dahmke, A.; Ebert, M.; Schafer, D.; Dethlefsen, F.
agents for bioremediation of nitrate contamination. Appl. Environ. Suitability of existing numerical model codes and thermodynamic
Microbiol. 1994, 60 (6), 1949−1955. databases for the prognosis of calcite dissolution processes in near-
(56) Liessens, J.; Vanbrabant, J.; Devos, P.; Kersters, K.; Verstraete, surface sediments due to a CO2 leakage investigated by column
W. Mixed culture hydrogenotrophic nitrate reduction in drinking- experiments. Aquat. Geochem. 2014, 20 (6), 639−661.
water. Microb. Ecol. 1992, 24 (3), 271−290. (77) Berta, M.; Dethlefsen, F.; Ebert, M.; Gundske, K.; Dahmke, A.
(57) Xia, S. Q.; Zhong, F. H.; Zhang, Y. H.; Li, H. X.; Yang, X. Bio- Surface passivation model explains pyrite oxidation kinetics in column
reduction of nitrate from groundwater using a hydrogen-based experiments with up to 11 bar p(O2). Environ. Earth Sci. 2016, 75,
membrane biofilm reactor. J. Environ. Sci. 2010, 22 (2), 257−262. 1175.
(58) Nevatalo, L. M.; Bijmans, M. F. M.; Lens, P. N. L.; Kaksonen, A. (78) Berger, S. Laborversuche zur Quantifizierung von Wasserstoff-
H.; Puhakka, J. A. Hydrogenotrophic sulfate reduction in a gas-lift Oxidationsreaktionen in Aquifer-Sedimenten. Kiel University, Kiel, 2015.
bioreactor operated at 9 degrees C. J. Microbiol. Biotechnol. 2010, 20 (79) Heron, G.; Crouzet, C.; Bourg, A. C. M.; Christensen, T. H.
(3), 615−621. Speciation of Fe(II) and Fe(III) in contaminated aquifer sediments
(59) van Houten, B. H.; Meulepas, R. J.; van Doesburg, W.; Smidt, using chemical extraction techniques. Environ. Sci. Technol. 1994, 28
H.; Muyzer, G.; Stams, A. J. Desulfovibrio paquesii sp. nov., a (9), 1698−1705.
hydrogenotrophic sulfate-reducing bacterium isolated from a syn- (80) Canfield, D.; Raiswell, R.; Westrich, J. T.; Reaves, C.; Berner, R.
thesis-gas-fed bioreactor treating zinc- and sulfate-rich wastewater. Int. A. The use of Chromium reduction in the analysis of reduced
J. Syst. Evol. Microbiol. 2009, 59 (Pt 2), 229−233.
inorganic sulfur in sediments and shales. Chem. Geol. 1986, 54, 149−
(60) Pfeiffer, W. T.; Beyer, C.; Bauer, S. Hydrogen storage in a
155.
heterogeneous sandstone formation: dimensioning and induced
(81) Appelo, C.; Postma, D. Geochemistry, groundwater and pollution,
hydraulic effects. Pet. Geosci. 2017, 23, 315.
(61) Feldmann, F.; Hagemann, B.; Ganzer, L.; Panfilov, M. Second ed.; Taylor & Francis: Leiden, 2005.
(82) Parkhurst, D. L.; Appelo, C. A. J. Description of input and
Numerical simulation of hydrodynamic and gas mixing processes in
underground hydrogen storages. Environ. Earth Sci. 2016, 75 (16), examples for PHREEQC version 3A computer program for
1165. speciation, batch-reaction, one-dimensional transport, and inverse
(62) Sainz-Garcia, A.; Abarca, E.; Rubi, V.; Grandia, F. Assessment of geochemical calculations. In U.S. Geological Survey Techniques and
feasible strategies for seasonal underground hydrogen storage in a Methods, book 6, A43; USGS: Reston, VA, 2013; p. 497.
saline aquifer. Int. J. Hydrogen Energy 2017, 42 (26), 16657−16666. (83) Lever, M. A. Acetogenesis in the energy-starved deep biosphere
(63) Lovley, D. R. Minimum threshold for hydrogen metabolism in − a paradox? Front. Microbiol. 2012, 2, 284.
methanogenic Bacteria. Appl. Environ. Microbiol. 1985, 49 (6), 1530− (84) Cai, M. L.; Liu, J. X.; Wei, Y. S. Enhanced biohydrogen
1531. production from sewage sludge with alkaline pretreatment. Environ.
(64) Mascus, C. Laboratory experiments evaluating geochemical Sci. Technol. 2004, 38 (11), 3195−3202.
reactions in aquifers due to hydrogen intrusion. Christian-Albrechts- (85) Yuan, H. Y.; Chen, Y. G.; Zhang, H. X.; Jiang, S.; Zhou, Q.; Gu,
Universität zu Kiel, Kiel, 2015. G. W. Improved bioproduction of short-chain fatty acids (SCFAs)
(65) Oren, A. The bioenergetic basis for the decrease in metabolic from excess sludge under alkaline conditions. Environ. Sci. Technol.
diversity at increasing salt concentrations: implications for the 2006, 40 (6), 2025−2029.

L DOI: 10.1021/acs.est.7b05467
Environ. Sci. Technol. XXXX, XXX, XXX−XXX
Environmental Science & Technology Article

(86) Jesußek, A.; Köber, R.; Grandel, S.; Dahmke, A. Aquifer heat
storage: sulphate reduction with acetate at increased temperatures.
Environ. Earth Sci. 2013, 69 (5), 1763−1771.
(87) Metzgen, A.; Berta, M.; Dethlefsen, F.; Ebert, M.; Dahmke, A.
Impact of pH on hydrogen oxidizing redox processes in aquifers due to
gas intrusions. In EGU General Assembly, Vienna, Austria, April 23−28,
2017.
(88) Schäfer, D.; Schäfer, W.; Kinzelbach, W. Simulation of reactive
processes related to biodegradation in aquifers: 1. Structure of the
three-dimensional reactive transport model. J. Contam. Hydrol. 1998,
31 (1−2), 167−186.
(89) Schäfer, D.; Hornbruch, G.; Schlenz, B.; Dahmke, A.
Contaminant spreading assuming different kinetic approaches to
simulate microbial degradation. Grundwasser 2007, 12 (1), 15−25.
(90) Krumholz, L. R.; Harris, S. H.; Tay, S. T.; Suflita, J. M.
Characterization of two subsurface H2-utilizing bacteria, Desulfomi-
crobium hypogeium sp. nov. and Acetobacterium psammolithicum sp. nov.,
and their ecological roles. Appl. Environ. Microbiol. 1999, 65 (6),
2300−2306, http://aem.asm.org/content/65/6/2300.long
(91) Vallero, M. V. G.; Lettinga, G.; Lens, P. N. L. High rate sulfate
reduction in a submerged anaerobic membrane bioreactor (SAMBaR)
at high salinity. J. Membr. Sci. 2005, 253 (1−2), 217−232.

M DOI: 10.1021/acs.est.7b05467
Environ. Sci. Technol. XXXX, XXX, XXX−XXX

View publication stats

You might also like