You are on page 1of 10

Chemical Geology 438 (2016) 1–10

Contents lists available at ScienceDirect

Chemical Geology

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

Biological reduction of structural Fe(III) in smectites by a marine


bacterium at 0.1 and 20 MPa
Deng Liu a,⁎, Fengping Wang b, Hailiang Dong c,d, Hongmei Wang a, Linduo Zhao d, Liuqin Huang a, Lingling Wu e
a
State Key Laboratory of Biogeology and Environmental Geology, China University of Geosciences, Wuhan 430074, China
b
State Key Laboratory of Microbial Metabolism and State Key Laboratory of Ocean Engineering, Shanghai Jiao Tong University, Shanghai 200240, China
c
State Key Laboratory of Biogeology and Environmental Geology, China University of Geosciences, Beijing 100083, China
d
Department of Geology and Environmental Earth Science, Miami University, Oxford, OH 45056, USA
e
Department of Earth and Environmental Sciences, University of Waterloo, Waterloo, Ontario N2L 3G1, Canada

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

Article history: Microbial iron reduction has been implicated as an important biochemical reaction on Earth. The influence of en-
Received 15 December 2015 vironmental parameters (e.g., pressure) on bioreduction of ferruginous clay minerals, however, is not well
Received in revised form 20 May 2016 constrained. The objective of this study was to investigate microbial reduction of structural Fe(III) in smectite
Accepted 21 May 2016
minerals and associated mineral transformations under elevated hydrostatic pressure. Bioreduction experiments
Available online 25 May 2016
were performed in a hydrostatic pressure system at 0.1 and 20 MPa, in which lactate as the sole electron donor,
Keywords:
two smectites having different iron contents (montmorillonite SWy-2 and nontronite NAu-2) as the sole electron
Iron reduction acceptor, and a marine bacterium Shewanella piezotolerans strain WP3 as the reaction mediator with and without
Smectite illitization an electron shuttle anthraquinone-2,6-disulfonate (AQDS). Our results indicated that S. piezotolerans strain WP3
Hydrostatic pressure was capable of reducing structural Fe(III) in smectites, and AQDS enhanced the initial reduction rate and final ex-
Shewanella tent. In the absence of AQDS, inhibitory effect of hydrostatic pressure on smectite reduction was observed. How-
ever, with AQDS, the reduction extents at 0.1 and 20 MPa were approximately the same value, but the initial
reduction rate at 20 MPa was lower than that at 0.1 MPa. Greater degree of smectite dissolution was found in
NAu-2 reactors compared to SWy-2 experiments. Mineralogical analysis by X-ray diffraction (XRD), Sybilla sim-
ulation, and scanning and transmission electron microscopy (SEM and TEM) showed that neoformation of illite
was present in bioreduced NAu-2, but not in SWy-2, and this smectite illitization was facilitated by higher hydro-
static pressure. These results have important implications for understanding iron cycling and low-temperature
illite formation in marine settings.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction nature (Favre et al., 2006; Stucki et al., 2007). Virtually all 2:1 type
clay minerals, including the classes of smectite, chlorite, vermiculite, il-
Iron (Fe) is the fourth most abundant element in Earth's crust, and lite and mica, contain ferric Fe in various amounts (Stucki et al., 1988).
comprises approximately 3.9% of elemental mass in sedimentary envi- For example, smectite, one of the largest classes of clay minerals, con-
ronments (Shelobolina et al., 2004). Iron redox cycling between its oxi- tains significant amounts of iron in its structure, ranging from
dized and reduced forms regulates a number of environmental 0.4 mmol Fe(III)/g for Wyoming Na-montmorillonite (SWy-1) to
processes, such as nutrient cycling (Roden and Edmonds, 1997), con- 4.2 mmol Fe(III)/g for nontronite (NAu-2) (Zhang et al., 2007a). Struc-
taminant migration and retention (Borch et al., 2010), and organic car- turally-coordinated Fe(III) in different clay minerals has been shown
bon preservation and mineralization (Taylor and Konhauser, 2011; as an electron acceptor for microbial respiration (Dong et al., 2009,
Lalonde et al., 2012). Microorganisms have been recognized as impor- and references therein). Previous studies have also demonstrated that
tant agents in mediating the Fe redox transition (Lovley, 1991; Weber a wide variety of anaerobic microorganisms (e.g., dissimilatory Fe(III)-
et al., 2006; Konhauser et al., 2011), especially concerning solid-phase reducing bacteria (DIRB): Kostka et al., 1996; Kim et al., 2004; Jaisi et
Fe reduction and oxidation (Melton et al., 2014). al., 2005; Zhang et al., 2007a; sulfate-reducing bacteria (SRB): Li et al.,
Given the ubiquitous occurrence of clay minerals in Earth surficial 2004; Liu et al., 2012; and some archaea: Liu et al., 2011; Zhang et al.,
environments and the occupancy of considerable Fe in their crystal 2012, 2013), are capable of reducing structural Fe(III) in clays.
structures, clay minerals are thought of as one of the main Fe pools in It is well established that the physical and chemical properties of
clay minerals could be significantly changed upon microbial Fe reduc-
⁎ Corresponding author. tion (Stucki and Kostka, 2006). For instance, as a result of Fe(III) reduc-
E-mail address: liud_cug@126.com (D. Liu). tion to Fe(II), specific surface area of clay minerals decreases, cation

http://dx.doi.org/10.1016/j.chemgeo.2016.05.020
0009-2541/© 2016 Elsevier B.V. All rights reserved.
2 D. Liu et al. / Chemical Geology 438 (2016) 1–10

exchange capacity increases, and swelling in water decreases (Stucki growth at optimal temperature of 15–20 °C, and with a broad pressure
and Kostka, 2006). In addition to aforementioned alteration of mineral optimum from 0.1 to 20 MPa (Xiao et al., 2007; Wang et al., 2008; Wu et
structure, some biogenic minerals could form associated with reductive al., 2013). Prior to bioreduction experiments, S. piezotolerans WP3 was
dissolution of pre-existing clay minerals, such as illite (Kim et al., 2004; cultured aerobically in marine 2216E medium (5 g/L tryptone, 1 g/L
Zhang et al., 2007a,b; Jaisi et al., 2011; Koo et al., 2014; Liu et al., 2012, yeast extract, 34 g/L NaCl, pH = 7.8) at 20 °C and one atmospheric pres-
2014, 2015), amorphous silica (Dong et al., 2003; O'Reilly et al., 2005; sure. Harvested WP3 cells in the mid to late log phase were washed
Zhang et al., 2007a; Liu et al., 2011), calcite (Li et al., 2004; Jaisi et al., three times with anoxic, modified marine salt bicarbonate buffer (dis-
2011), and other secondary minerals (e.g., vivianite, siderite) (Dong et tilled water 1 L, NaHCO3 2.5 g, NaCl 35 g, MgCl2·6H2O 2 g, KCl 1 g and
al., 2009). Among these byproducts, low-temperature illite that forms CaCl2 0.1 g, pH = 7.8), and then re-suspended in the aforementioned
through the bioreduction of smectite has attracted much attention, anaerobic buffer to obtain a cell concentration of ~2 × 108 cells/mL.
mainly because the formation of illite facilitated by microbes at ambient
temperature challenges our conventional wisdom that illite formation is 2.3. Bioreduction experiments
an abiotic process that typically requires conditions of 300–400 °C,
100 MPa, and 4–5 months (Kim et al., 2004; Dong et al., 2009; Dong, To evaluate the effect of pressure on microbial reduction activity and
2012). mineral transformation, the bioreduction reactors with strain WP3
Even though our knowledge of microbe-clay mineral interaction has were incubated under 0.1 MPa or 20 MPa. SWy-2 or NAu-2 minerals
been rapidly expanding, it is worthwhile to note that previous studies were suspended (final concentration 5 g/L) with a modified marine
on bioreduction of Fe(III) in clay minerals have used microbes that salt bicarbonate buffer in 120-mL serum bottles. After adjusting pH to
were originally isolated from terrestrial environments (e.g., Shewanella 7.8 with 1 M NaOH, the bottles were purged with ultra-pure N2, sealed
oneidensis MR-1 from freshwater sediments of Lake Oneida, S. with thick butyl rubber stoppers, and capped with aluminum caps. After
putrefaciens CN32 from shale-sandstone core samples). Studies on mi- the bottles were autoclaved, cells from an anaerobic stock were added
crobial Fe(III) reduction in clay minerals under marine conditions and into the bottles to achieve a final concentration of ~ 2 × 107 cells/mL.
associated mineral transformations are rare. The deep-sea environ- Lactate (electron donor) from stock solution was injected into the bot-
ments, as we know, are subjected to highly elevated hydrostatic pres- tles to achieve a final concentration of 20 mM. A quantity of 0.1 mM of
sure, and thus are unique habitats for piezotolerant/piezophilic iron- anthraquinone-2,6-disulfonate (AQDS; electron shuttle) was added in
reducing microorganisms. Clay minerals are generally considered as selected bottles to facilitate electron transfer. In an anaerobic chamber
the most abundant mineral constituents of marine sediments (Griffin (filled with 98% N2 and 2% H2, Coy Laboratory Products, Michigan),
et al., 1968). However, the question whether hydrostatic pressure influ- the well-mixed experimental suspensions (mineral, cells, lactate, buffer
ences microbial reduction of clay Fe(III) and the consequent mineral without/with AQDS) were immediately dispensed into 10-mL sterile in-
transformations (e.g., smectite-to-illite reaction) has not been ad- jection syringes and placed in steel pressure vessels. The pressure ves-
dressed. Therefore, the objective of this study was to investigate smec- sels were pressurized by injecting water through a hydrostatic
tite reduction by a marine DIRB at two different hydrostatic pressures pressure system (HPS). The composition and operation details of HPS
(0.1 and 20 MPa, corresponding to 1 and 200 atmospheric pressure, re- were described elsewhere (Wu et al., 2013). Once at the required pres-
spectively) to improve our understanding of biogeochemical Fe cycling sure, the pressure vessels were then placed in an incubator at 20 °C.
in marine systems.
2.4. Chemical analyses
2. Materials and methods
The total Fe(II) content was measured by the 1,10-phenanthroline
2.1. Smectite preparation assay (Amonette and Templeton, 1998). The 1,10-phenanthroline
method is based on total digestion of the mineral with HF-H2SO4 and
Two smectites were selected: Wyoming montmorillonite SWy-2 has been shown to be an effective approach in measuring total Fe(II)
and nontronite NAu-2. These two samples have a similar structure, in phyllosilicates (Anastacio et al., 2008). The initial rate and extent of
but different iron content. Both SWy-2 and NAu-2 bulk materials were Fe(III) bioreduction were calculated by following equations:
purchased from the Source Clays Repository of the Clay Minerals Society
(http://www.clays.org/SOURCE%20CLAYS/SCdata.html). Bulk smectite Initial Reduction rate 
samples were manually ground with an agate mortar and soaked in FeðIIÞfinal −FeðIIÞinitial within the first 24 h
¼ ð1Þ
0.5 M NaCl solution overnight. The 0.02–0.5 μm size fraction of smectite 24 h
samples was subsequently collected by centrifugation using Stokes' 
settling law. The chloride anion was removed by repeated washing FeðIIÞtotal −FeðIIÞinitial
Reduction extent ¼  100%: ð2Þ
in doubly distilled water (ddH2O) and its complete removal was FeðIIIÞtotal
confirmed by AgNO3 titration. Smectites were then air-dried and
used for experiments. Previous studies have shown that SWy-2 To monitor reductive dissolution of smectites, concentrations of sol-
contains 2.3 wt% Fe, of which 97.1% is Fe(III), and its formula can be uble Fe and Si were measured over the course of the experiments. A 1-
written as (Ca0.16Na0.24)(Si6.73Al1.27)(Al1.45Fe0.13Mg0.44)O20(OH)4 mL aliquot was sampled inside an anaerobic chamber, centrifugated
(Bishop et al., 2011); and NAu-2 contains 23.4 wt% Fe, of (10,000 × g, 10 min), and then passed through a 0.45-μm filter into a
which 99.4% is Fe(III), and its structural formula can be expressed as 5-mL tube. The acidified filtrate samples were analyzed by inductively
M0.72(Si7.55Al0.16Fe0.29)(Al0.34Fe3.54Mg0.05)O20(OH)4 (M represents the coupled plasma optical emission spectrophotometry (ICP-OES) (ICAP-
interlayer cation) (Keeking et al., 2000; Jaisi et al., 2005). 6300, ThermoFisher Scientific, USA).

2.2. Bacterial strain and culture medium 2.5. X-ray diffraction (XRD)

Shewanella piezotolerans strain WP3, a gram-negative facultative Smear mounts (Moore and Reynolds, 1997) were prepared for all
bacterium, was isolated from west Pacific deep-sea sediment located XRD analyses of both control and bioreduced smectites after 42-days
at a water depth of 1914 m (Xiao et al., 2007). Physiological and bio- of incubation to identify any mineralogical changes. Clay mineral slur-
chemical tests have shown that strain WP3 is able to grow on various ries of 0.5 mL in volume were smeared onto glass slides and dried at
substrates such as nitrate, Fe3+, fumarate and dimethyl sulfoxide for 30 °C overnight in an anaerobic chamber. To identify illitic layers, air-
D. Liu et al. / Chemical Geology 438 (2016) 1–10 3

dried and oriented clay slides were saturated with ethylene glycol vapor 3. Results
overnight at 60 °C to expand smectite interlayers. XRD patterns of eth-
ylene glycolated samples were collected by a Scintag X-ray powder dif- 3.1. Bioreduction of Fe(III) in SWy-2 and NAu-2 at 0.1 and 20 MPa
fractometer using wavelength of Cu Kα, a fixed slit scintillation detector,
and power of 14,000 W (40 kV, 35 mA). The XRD slides were scanned No appreciable Fe(II) accumulation was detected in abiotic controls
from 3 to 12° 2θ stepping at 0.02° with a count time of 2 s per step. either at 0.1 or 20 MPa (Fig. 1). By contrast, Fe(II) concentration steadily
The XRD patterns of ethylene glycolated samples were modeled increased with time within the inoculated experimental bottles, indicat-
using the Sybilla program (McCarty et al., 2009) to quantify the relative ing that strain WP3 was capable of reducing structural Fe(III) in clay
proportions of smectite, mixed-layer illite-smectite (I-S) and discrete il- minerals (Fig. 1). Similar patterns of Fe(III) reduction were observed
lite in bioreduced samples. in the bioreactors with SWy-2 and NAu-2: without AQDS, higher pres-
sure (20 MPa) decreased the rate of Fe(II) production (Fig. 1A & B); in
the treatments with AQDS, however, the concentrations of produced
2.6. Scanning and transmission electron microscopy (SEM and TEM) Fe(II) within the first 7 days were only slightly lower at 20 MPa than
at 0.1 MPa, and gradually reached the similar levels afterwards
The spatial associations between smectites and cells of strain WP3, (Fig. 1C & D). At the end of microbial reduction, there was much less
and the morphology and particle size of biogenic minerals were Fe(II) accumulation in all SWy-2 setups, compared to those reacted
examined by SEM. Cells-mineral suspensions were fixed with 2% with NAu-2: depending on the presence/absence of AQDS, 0.060 to
paraformaldehyde and 2.5% glutaraldehyde. After fixation, 0.2-mL 0.154 mmol/g Fe(II) were detected in SWy-2 bioreactors (Fig. 1A & C),
suspension was dropped onto the surface of a glass cover slip. The whereas 0.396 to 1.162 mmol/g in the NAu-2 bottles (Fig. 1B & D).
sample-coated cover slip was sequentially dehydrated using varying The addition of AQDS greatly stimulated the bioreduction of Fe(III) ei-
proportions of ethanol followed by critical point drying with a ther at 0.1 or 20 MPa, probably via enhancement of electron transfer
Tousimis Samdro-780A Critical Point Dryer (CPD) (Dong et al., (Dong et al., 2003; Jaisi et al., 2005).
2003). The cover slip was mounted onto a SEM stub and coated The initial reduction rate by WP3 was calculated to assess the effect
with Au. The sample was then analyzed using a Zeiss Supra 35 VP of hydrostatic pressure on bioreduction. In the SWy-2 experiments, the
SEM with Genesis 200 X-ray energy dispersive spectroscopy (EDS). initial rate at 0.1 MPa was 1.00 and 4.94 μmol/g/h without and with
The SEM was operated at an accelerating voltage of 10–15 keV. A AQDS, respectively; whereas the rate was 0.66 and 3.58 μmol/g/h with-
short working distance (6–10 mm) and low beam current (30–40 μA) out and with AQDS at 20 MPa. In comparison, the initial rate of
were used to achieve the best image resolution. bioreduction for NAu-2 at 0.1 MPa was 2.76 and 9.84 μmol/g/h without
Bioreduction-induced mineralogical changes were further studied and with AQDS, respectively, significantly higher than those at 20 MPa
by TEM. Diluted suspensions were pipetted onto 300 mesh copper (1.34–6.92 μmol/g/h).
grids with a nitrocellulous membrane and carbon coating. The grids At the end of 42 days, the extent of Fe(III) bioreduction varied for dif-
were prepared and allowed to dry in an anaerobic chamber. A JEOL ferent smectites and was enhanced by the presence of AQDS (Fig. 2).
JEM-2100 LaB6 TEM with a 200 keV accelerating voltage was used for Specifically, in the absence of AQDS, the reduction extent in SWy-2
TEM analysis. EDS fitted in the TEM was employed for mineral reached 19.3% and 15.1% under the pressure conditions of 0.1 and
identification. 20 MPa, respectively; whereas the reduction extent in SWy-2 was

Fig. 1. Smectite reduction by WP3 under hydrostatic pressures of 0.1 and 20 MPa. Fe(II) increased with time in montmorillonite SWy-2 (A, C) and nontronite NAu-2 (B, D) reactors due to
bioreduction in the treatments without and with AQDS. All results were from duplicate cultures and the error bars represent two-sigma variation.
4 D. Liu et al. / Chemical Geology 438 (2016) 1–10

SWy-2 systems (Fig. 3A), but to 119 μM (0.1 MPa) and 97 μM (20 MPa)
in NAu-2 experiments (Fig. 3B). In the presence of AQDS, the final con-
centration of soluble Fe rose to 14 μM (0.1 MPa) and 13 μM (20 MPa) in
SWy-2 reactors (Fig. 3A), but to 170 μM (0.1 MPa) and 139 μM (20 MPa)
in NAu-2 reactors (Fig. 3B).
The soluble Si concentration in biotic reactors exhibited a similar
trend to those of aqueous Fe (Fig. 4). By day 42, the concentrations of
soluble Si in NAu-2 ranged from 169 to 797 μM (depending on the ab-
sence/presence of AQDS), remarkably higher than those in SWy-2 reac-
tors (ranging from 27 to 55 μM).

3.3. XRD and related modelling

Experimental XRD patterns of glycolated oriented mounts are


shown in Fig. 5. After bioreduction, the 001 peak of smectites had a
Fig. 2. A comparison among various reduction extents with four different treatments for
each smectite mineral. broader distribution and a lower intensity, indicating that smectites
partially dissolved during the experiments. In comparison with SWy-2
(Fig. 5A & B), a small hump in the 8° to 11° 2θ (corresponding to d values
enhanced by AQDS up to about 38.2% at both 0.1 and 20 MPa. For NAu-2, ranging from 0.804 to 1.104 nm) emerged in bioreduced, AQDS-
14.0% (0.1 MPa) and 9.5% (20 MPa) of Fe(III) were reduced in the ab- amended NAu-2 samples (Fig. 5C & D). According to our previous stud-
sence of AQDS; the reduction extents under two pressure conditions, ies (e.g., Liu et al., 2014, 2015), this hump might be an overlapping peak
however, had approximately same value of 27.5% with AQDS. including smectite, mixed-layer I-S, and discrete illite.
In summary, the above data indicated that (i) AQDS could enhance To verify the presence of illite, the XRD patterns of bioreduced NAu-
the rate and extent of microbial reduction of Fe(III) in smectites; (ii) 2 with AQDS were further modeled by the Sybilla program. This simula-
high pressure slowed down the initial rate of smectite reduction; (iii) tion using curve decomposition method can provide precise informa-
high pressure decreased the final extent of smectite reduction when tion on peak positions and areas. The modelling results showed that
AQDS was absent, whereas the presence of AQDS alleviated the inhibi- neoformations of R0 I-S and illite occurred upon bioreduction of NAu-
tory effect of elevated pressure on Fe(III) bioreduction. 2 (Fig. 6). It is interesting to note that more illite formed in the treat-
ment of higher hydrostatic pressure (Fig. 6).
3.2. Fe and Si release from smectites during bioreduction of Fe(III)
3.4. SEM observation
Compared to negligible concentrations of soluble Fe in abiotic con-
trols (data not shown), aqueous Fe concentration in all biotic treatments Individual cells (Fig. 7A) and cell assemblages (Fig. 7B) were ob-
increased within 14–28 days and then reached a plateau (Fig. 3). The served under SEM in biotic systems after two-week incubation. Interest-
final concentration of aqueous Fe in all biotic reactors with NAu-2 was ingly, cell assemblages generally attached to smectite particle edges
approximately 10 times higher than those with SWy-2. Specifically, (Fig. 7C). The magnified views (Fig. 7D) showed exopolymeric sub-
without AQDS, the concentration of soluble Fe by the end of the exper- stances (EPS)-like and nanowire (pili)-like materials connecting cells
iments (42 days) increased to 10 μM (0.1 MPa) and 11 μM (20 MPa) in and smectite.

Fig. 3. Aqueous Fe production with time in different bioreactors: (A) SWy-2 without AQDS; (B) NAu-2 without AQDS; (C) SWy-2 with AQDS; (D) NAu-2 with AQDS.
D. Liu et al. / Chemical Geology 438 (2016) 1–10 5

Fig. 4. Aqueous Si production with time in different bioreactors: (A) SWy-2 without AQDS; (B) NAu-2 without AQDS; (C) SWy-2 with AQDS; (D) NAu-2 with AQDS.

Fig. 5. XRD patterns of ethylene glycolated bioreduced smectite clays under 0.1 and 20 MPa (Sme, smectite; Ill, illite). (A) Unreduced and reduced SWy-2 at 0.1 MPa; (B) unreduced and
reduced SWy-2 at 20 MPa; (C) unreduced and reduced NAu-2 at 0.1 MPa; (D) unreduced and reduced NAu-2 at 20 MPa. The dash line in each panel indicates the position of illite (001)
peak. The black arrows shown in bottom panels indicate a possible illite peak.
6 D. Liu et al. / Chemical Geology 438 (2016) 1–10

Fig. 6. Comparison of the experimental XRD patterns from bioreduced NAu-2 (solid curves) with calculated patterns (dash curves) based on Sybilla simulation: (A) at 0.1 MPa; (B) at
20 MPa. The Sybilla modelling results show the formation of R0 I-S and discrete illite after bioreduction under hydrostatic pressures of 0.1 MPa (C) and 20 MPa (D).

Residual solid phases were also collected at the end of bioreduction (Fig. 10). TEM results indicated that the (001) fringes of unreduced
experiments for SEM observations. SWy-2 particles in both abiotic con- NAu-2 were anastomosing and wavy with a 1.2–1.3 nm layer spacing,
trol and bioreduced sample did not undergo any mineralogical changes, a low Al/Si ratio, and a high amount of Fe (Fig. 10A). In bioreduced sam-
and were characteristic of the typical flaky shape (Fig. 8). Compared to ples, nontronite aggregations appeared as crumpled flakes with rolled
abiotic control (Fig. 9A), however, various altered and secondary min- edges (Fig. 10B). In addition, illite packets having a higher Al/Si ratio
erals in blocky or platy shapes appeared in bioreduced NAu-2 (Fig. and K content were found after bioreduction. The fringes of illite were
9B). The SEM-EDS results revealed that the byproducts consisted of relatively straight and had a constant layer spacing of 1.0 nm (Fig. 10C).
euhedral smectite (particle a in Fig. 9C), spherical silica (particle b in
Fig. 9C), platy illite particles (particles c & d in Fig. 9C), nontronite
aggregates (Fig. 9D), and rhombohedral calcite crystal in association 4. Discussion
with silica (Fig. 9E).
4.1. Rate and extent of smectite reduction by strain WP3 in comparison with
3.5. TEM observation other Shewanella species

The bioreduced, AQDS-amended NAu-2 collected from 20 MPa reac- Recently, a growing body of experimental studies has shown that
tors was selected as a representative to further verify illite formation strains in Shewanella genus are capable of reducing Fe(III) in smectite

Fig. 7. SEM images indicate the association of WP3 cells and NAu-2 particles. Two different associations were observed: individual cell (A) and cell assemblages (B). The cell assemblages
mostly attached on the edge of NAu-2 particle (C). The magnified views (D and E) show the pili- and EPS-like substances.
D. Liu et al. / Chemical Geology 438 (2016) 1–10 7

(oxyhydr)oxides, clay minerals usually carry a net negative charge


(Sposito, 1984). Microbial surface is also commonly negatively
charged under growth pH conditions, thereby resulting in a repul-
sive force between cells and clay minerals. It has been demonstrated
that, however, positive charges could possibly exist at edge sites of
clay minerals (Sposito, 1984). Hence, iron-reducing microorganisms,
like WP3, preferably attach onto clay edges for respiring structurally-
coordinated Fe(III) (Dong et al., 2009). Such microbe-clay mineral
association leads to the fact that much of the structural Fe(III) in
smectite may not be accessible to iron-reducing microorganisms.
However, the incomplete Fe reduction was still observed when AQDS
was amended into bioreactors. The above case might be ascribed to
the toxic and electron-flow blocking effects of released cations from
reductive dissolution of smectite. Indeed, an experimental study by
Jaisi et al. (2007) revealed that aqueous Fe(II) is toxic to cells and if
sorbed onto the surface of clay minerals and microbial cells, Fe(II)
imposed a thermodynamic “barrier” on electron transfer from DIRB to
mineral surfaces. Furthermore, the toxicity of Al3+ to iron-reducing mi-
croorganisms has been recognized recently (Williams and Hillier, 2014;
Liu et al., 2016).

4.2. Influence of hydrostatic pressure on smectite reduction by strain WP3

Our results indicated that hydrostatic pressure decreased the


smectite reduction rate and extent by strain WP3 without exogenous
electron transfer mediators. The inhibitory effect of hydrostatic pres-
sure on bioreduction of Fe(III) was also observed with the same
strain when it used ferrihydrite as electron acceptor (Wu et al.,
Fig. 8. SEM images of SWy-2 samples after 42-days of incubation: (A) abiotic control; (B)
bioreduced SWy-2. 2013). In the aforementioned work, the bioreduction extent of
ferrihydrite at 20 MPa decreased approximately 30% of what it was
at 0.1 MPa, similar to the results of our study using either SWy-2
(~ 22% decline) or NAu-2 (~ 32% decline) as electron acceptor. In
clays, for instance, S. oneidensis MR-1 (Kostka et al., 1996; Dong et al., addition to piezotolerant strain WP3, ferric citrate reduction experi-
2003; Kim et al., 2004; O'Reilly et al., 2005), S. putrefaciens CN32 (Jaisi ments using piezophilic DIRB S. profunda LT13a also supported the
et al., 2005, 2011; Zhang et al., 2007a; Bishop et al., 2011) and S. algae above findings (Picard et al., 2015).
BrY (Liu et al., 2014). In the present study, S. piezotolerans WP3 isolated There are at least two explanations for the inhibition of smectite re-
from deep-sea sediment was also able to respire structural Fe(III) in duction by hydrostatic pressure. First, it has been hypothesized that the
smectites. In addition, our results indicated that the rate and extent of kinetics of enzyme-involving reactions under high pressure are
Fe(III) reduction by WP3 varied with smectite type, addition of AQDS, governed by Le Chatelier's principle, that is, high pressure favors the re-
and hydrostatic pressure (Figs. 1 & 2). actions with a net decrease in volume of reactants, while the reactions
Because there are few data on SWy-2 reduction by Shewanella in having a net increase in volume are retarded (Balny et al., 1997; Wu
the literature, we compared the rate and extent of NAu-2 reduction et al., 2013). Considering the experiments here, microbial respiration
by WP3 with other well-known Shewanella species under similar ex- of structural Fe(III) in smectites is accompanied by a volume increase
perimental conditions, such as 5 g/L NAu-2, non-growth medium, due to CO2 production, according to the following reaction:
ambient pressure and inoculum density of 107–108 cells/mL. In the 4Fe3 + + C3H5O− 3 (lactate) + H2O → 4Fe2 + + C2H3O− 2
absence of AQDS, the initial rate by WP3 was 2.76 μmol/g/h at (acetate) + CO2 + 4H+. As such, smectite reduction by WP3 is ther-
1 atm, very close to that of NAu-2 by S. algae BrY (3.20 μmol/g/h) modynamically inhibited under elevated hydrostatic pressures. In
(Liu et al., 2014), but nearly twice the rate of NAu-2 by S. putrefaciens addition, bioavailability of structural Fe(III) in smectites might
CN32 (1.60 μmol/g/h) (Jaisi et al., 2005). Once AQDS was added as change under high pressure due to the alteration of mineralogical
an electron shuttle, the initial rate by WP3 was enhanced to property, and thus may influence microbial reduction efficiency.
9.84 μmol/g/h, which was significantly lower than that by either For instance, it has been shown that smectite swelling decreased
CN32 (30.40 μmol/g/h) (Jaisi et al., 2005) or BrY (56.80 μM/h) under elevated hydrostatic pressure, due to decreased thickness of
(Liu et al., 2014). The reduction extents by WP3, 14.0% and 27.5% smectite interlayer (Wang et al., 1980; Morin and Silva, 1984). In
without and with AQDS, were in the range of those achieved by doing so, a decreased osmotic swelling caused by higher pressure
other Shewanella strains (Jaisi et al., 2005; Zhang et al., 2007a; can lead to fewer edge sites for microbial attachment and subse-
Bishop et al., 2011; Liu et al., 2014). quently lower reduction extent (Liu et al., 2012).
Incomplete reduction of structural Fe(III) in smectite clays was ob- In the presence of AQDS, the final reduction extent at 20 MPa was
served not only for Shewanella, but also for other types of anaerobic mi- similar to that at 0.1 MPa (Fig. 2). As described before, direct contact
croorganisms, for example, SRB (Li et al., 2004; Liu et al., 2012) and is a major mode for microbial respiration on structurally-coordinat-
methanogens (Liu et al., 2011; Zhang et al., 2012, 2013). Several factors ed Fe(III) in smectites. Once exogenous electron shuttles, like AQDS
may account for such incomplete Fe reduction in clay minerals (e.g., used in this study, are added into the bioreactors, they can be rapidly
Dong et al., 2009), including electron pathway, energetics and elemen- reduced by Shewanella, and the reduced form AH 2DS can subse-
tal toxicity. In general, in the absence of any external electron shuttles quently reduce Fe(III) chemically (Lovley et al., 1996). In this case,
(e.g., AQDS), direct contact between cells and mineral surface is the electron shuttle-catalyzed Fe reduction pathway plays a more
generally believed to be a major strategy of electron transfer from cell important role than the direct cell-to-mineral contact pathway
membrane to the structural Fe(III) (Shi et al., 2007). Unlike iron (Lovley et al., 1996). Because of AQDS accessibility to both edge
8 D. Liu et al. / Chemical Geology 438 (2016) 1–10

Fig. 9. SEM images and SEM-EDS compositions show NAu-2 alteration and transformation after bioreduction at 0.1 MPa: (A) abiotic control showing flaky nontronite particles; (B)
bioreduced NAu-2 with AQDS showing plate-like byproduct and nontronite aggregates; (C) residual nontronite (grain a), silica globule (grain b) and illite (grains c and d) in
bioreduced NAu-2 samples; the occurrence of nontronite aggregates (D) and calcite-like particle (E) after bioreduction.

sites and the basal surface, electron transfer from electron shuttles to the reduction extent with AQDS is therefore enhanced relative to
structural Fe(III) in clay minerals could occur both parallel and per- that without AQDS, and is less affected by hydrostatic pressure
pendicular to the clay (001) layer (Dong et al., 2009). Therefore, (Figs. 1 & 2).

Fig. 10. TEM micrographs for unreduced and bioreduced NAu-2 with AQDS at 0.1 MPa: (A) unreduced NAu-2 displaying 1.3 or 1.2 lattice fringes; (B) nontronite aggregates; (C) illite packet
with 1.0 nm layer spacing. The insert in panel A is the EDS spectrum of NAu-2. The insert EDS in panel C shows a typical illite composition.
D. Liu et al. / Chemical Geology 438 (2016) 1–10 9

4.3. The smectite-to-illite reaction 5. Conclusions

Thesmectite-to-illitereaction(smectite+Al3+ +K+ →illite+silica), The piezotolerant, iron-reducing bacterium S. piezotolerans WP3,


silica), also termed smectite illitization, is one of the most important re- originally isolated from deep-sea sediment, was able to reduce structur-
actions in burial diagenesis as it greatly influences various geological al Fe(III) in two smectite clays (nontronite NAu-2 and montmorillonite
processes such as development of geopressure, petroleum evolution, SWy-2) under hydrostatic pressure conditions. The addition of AQDS
sediment porosity, and growth of faults (Hower et al., 1976; Pevear, significantly enhanced the smectite reduction by WP3. Without AQDS,
1999; Dong et al., 1997). This reaction is generally considered to pro- both initial bioreduction rate and final extent of smectite at 20 MPa
ceed through mixed-layer illite-smectite (I-S) intermediates [e.g., ran- were lower than those at 0.1 MPa. In the presence of AQDS, only initial
dom R0 I-S, ordered R1, R2, R3 I-S (R, Reichweite ordering parameter), bioreduction rate decreased under elevated pressure. During microbial
and end-member illite] in which the percentage of illite layers increases Fe reduction, a greater degree of mineral dissolution was observed in
with increasing temperature, geological time, pressure, pH, water/rock NAu-2, likely because of its higher Fe(III) content than SWy-2. A com-
ratio and potassium concentration (Dong et al., 2009, and references bined XRD, SEM, and TEM characterizations demonstrated that the
therein). In addition to above abiotic factors, some anaerobic microor- smectite-to-illite transformation occurred upon bioreduction of Fe(III)
ganisms, especially DIRB, have been recently proposed as considerable in NAu-2 but not in SWy-2 reactors, and this reaction was promoted
geo-catalysts in smectite illitization (Kim et al., 2004; Zhang et al., by elevated pressure.
2007a,b; Dong et al., 2009; Vorhies and Gaines, 2009; Jaisi et al., 2011;
Koo et al., 2014; Liu et al., 2012, 2014, 2015). According to these reports, Acknowledgments
microorganisms promote smectite illitization by a dissolution-precipi-
tation mechanism in which smectite dissolves upon microbial Fe reduc- This research is financially supported by grants from the National
tion and soluble constituents (e.g., Al, Si and Fe) recombine to Natural Science Foundation of China (Nos. 41302270, 41572328,
precipitate as illite if the solution contains sufficient K (Zhang et al., 41572323, 91228201 and 91428308), the 111 Project (No. B08030),
2007a,b; Jaisi et al., 2011; Liu et al., 2012, 2014, 2015). It is notable and the Fundamental Research Funds for the Central Universities
that the smectite end-member nontronite was typically used in previ- (CUGL140815 and CUGL150409). The JEOL 2100 TEM used in this
ous studies. In contrast to the occurrence of biogenic illite in bioreduced study was supported by NSF grant EAR-0722807. We are grateful to
nontronite samples, our present work further showed that the experi- Dr. Jeremy B. Fein, Dr. Joseph W. Stucki and an anonymous reviewer
ments with Fe-poor smectite SWy-2 did not produce any detectable il- whose comments improved the quality of this manuscript.
lite after bioreduction. The reason for this discrepancy between SWy-2
and NAu-2 might be caused by their degree of reductive dissolution.
References
Considering the fact that the concentrations of released elements in
SWy-2 reactors were nearly 10 times lower than those in NAu-2 exper- Amonette, J.E., Templeton, J.C., 1998. Improvements to the quantitative assay of
iments, it is reasonable to speculate that soluble constituents produced nonrefractory minerals for Fe(II) and total Fe using 1,10-phenanthroline. Clay Clay
Miner. 46 (1), 51–62.
during reduction of Fe(III) in SWy-2 did not reach the saturation level Anastacio, A.S., Harris, B., Yoo, H.I., Fabris, J.D., Stucki, J.W., 2008. Limitations of the
with respect to illite precipitation. ferrozine method for quantitative assay of mineral systems for ferrous and total
In the bioreduced NAu-2 samples with AQDS, XRD results suggested iron. Geochim. Cosmochim. Acta 72 (20), 5001–5008.
Balny, C., Mozhaev, V.V., Lange, R., 1997. Hydrostatic pressure and proteins: basic con-
that higher pressure promoted formation of more illitic layers (both in- cepts and new data. Comp. Biochem. Physiol. 116A (4), 299–304.
terstratified and discrete illite) despite a similar Fe(III) reduction extent Bishop, M.E., Dong, H., Kukkadapu, R.K., Liu, C., Edelmann, R.E., 2011. Bioreduction of Fe-
under two different pressures. A two-step mechanism has been pro- bearing clay minerals and their reactivity toward pertechnetate (Tc-99). Geochim.
Cosmochim. Acta 75, 5229–5246.
posed for the biogenic illite formation: reductive dissolution upon mi- Borch, T., Kretzschmar, R., Kappler, A., Van Cappellen, P., Ginder-Vogel, M., Voegelin, A.,
crobial activity and subsequent illite precipitation and ageing (Jaisi et Campbell, K., 2010. Biogeochemical redox processes and their impact on contaminant
al., 2011; Liu et al., 2014, 2015). It has been demonstrated that certain dynamics. Environ. Sci. Technol. 44 (1), 15–23.
Buatier, M.D., Peacor, D.R., O'Neil, J.R., 1992. Smectite-illite transition in Barbados accre-
physical parameters (e.g., pressure and temperature) and water chem-
tionary wedge sediments: TEM and AEM evidence for dissolution/crystallization at
istry (e.g., ionic strength, pH and, Al and K availability) typically control low temperature. Clay Clay Miner. 40 (1), 65–80.
the growth of illite crystal (Huang et al., 1993). In the present study, Dekov, V.M., Kamenov, G.D., Stummeyer, J., Thiry, M., Savelli, C., Shanks, W.C., Fortin, D.,
Kuzmann, E., Vértes, A., 2007. Hydrothermal nontronite formation at Eolo Seamount
with a similar reduction level, pressure might play a crucial role in accel-
(Aeolian volcanic arc, Tyrrhenian Sea). Chem. Geol. 245 (1–2), 103–119.
erating the rate of the second step, e.g., illite precipitation. However, to Dong, H., 2012. Clay-microbe interactions and implications for environmental mitigation.
understand the role of pressure in microbially-mediated smectite Elements 8, 113–118.
illitization, further experiments are warranted using a wider pressure Dong, H., Peacor, D.R., Freed, R.L., 1997. Phase relations among smectite, R1 I/S and illite.
Am. Mineral. 82, 379–391.
gradient. Dong, H., Kostka, J.E., Kim, J., 2003. Microscopic evidence for microbial dissolution smec-
tite. Clay Clay Miner. 51, 502–512.
4.4. Implications for biogeochemical transformations Dong, H., Jaisi, D.P., Kim, J.W., Zhang, G., 2009. Microbe-clay mineral interactions. Am.
Mineral. 94 (11−12), 1505–1519.
Favre, F., Stucki, J.W., Boivin, P., 2006. Redox properties of structural Fe in ferruginous
It is well known that Shewanella strains are the most abundant smectite. A discussion of the standard potential and its environmental implications.
Proteobacteria and, are believed to dominate the bacterial community Clay Clay Miner. 54, 466–472.
Griffin, J., Windom, H., Goldberg, E., 1968. The Distribution of Clay Minerals in the World
in deep-sea environment (Wang et al., 2008). Although nontronite is Ocean. Elsevier, Amsterdam.
not common among the smectite group minerals in soils and freshwater Hower, J., Eslinger, E.W., Hower, M.H., Perry, E.A., 1976. Mechanism of burial metamor-
settings, a growing body of work shows that massive nontronite can de- phism of argillaceous sediments. 1. Mineralogical and chemical evidence. Geol. Soc.
Am. Bull. 87, 725–737.
posit in certain areas of the deep sea, particularly seamounts such as Huang, W., Longo, J.M., Pevear, D.R., 1993. An experimentally derived kinetic model for
Juan de Fuca Ridge (Murnane and Clague, 1983), Eolo Seamount of the smectite-to-illite conversion and its use as a geothermometer. Clay Clay Miner. 41,
Tyrrhenian Sea (Dekov et al., 2007), and Valu Fa Ridge (Sun et al., 162–177.
Jaisi, D.P., Kukkadapu, R.K., Eberl, D.D., Dong, H., 2005. Control of Fe(III) site occupancy on
2012). Therefore, Shewanella and other DIRB can co-exist with
the rate and extent of microbial reduction of Fe(III) in nontronite. Geochim.
nontronite in deep-sea sediments. Present findings on the smectite Cosmochim. Acta 69, 5429–5440.
illitization catalyzed by marine DIRB have implications with regards to Jaisi, D.P., Dong, H., Liu, C., 2007. Influence of biogenic Fe(II) on the extent of microbial re-
the origin of low-temperature illite which was previously observed in duction of Fe(III) in clay minerals nontronite, illite and chlorite. Geochim.
Cosmochim. Acta 71, 1145–1158.
young marine sediments (Buatier et al., 1992; Vorhies and Gaines, Jaisi, D.P., Eberl, D.D., Dong, H., Kim, J., 2011. The formation of illite from nontronite by
2009). mesophilic and thermophilic bacterial reaction. Clay Clay Miner. 59, 21–33.
10 D. Liu et al. / Chemical Geology 438 (2016) 1–10

Keeking, J.L., Raven, M.D., Gates, W.P., 2000. Geology and characterization of two hydro- Roden, E.E., Edmonds, J.W., 1997. Phosphate moblization in iron-rich anaerobic sedi-
thermal nontronites from weathered metamorphic rocks at the Uley Graohite ments: Microbial Fe(III) oxide reduction versus iron-sulfide formation. Arch.
Mine, South Australia. Clay Clay Miner. 48, 537–548. Hydrobiol. 139, 347–378.
Kim, J., Dong, H., Seabaugh, J., Newell, S.W., Eberl, D.D., 2004. Role of microbes in the Shelobolina, E.S., Anderson, R.T., Vodyanitskii, Y.N., Sivtsov, A.V., Yuretich, R., Lovley, D.R.,
smectite-to-illite reaction. Science 303, 830–832. 2004. Importance of clay size minerals for Fe(III) respiration in a petroleum-contam-
Konhauser, K.O., Kappler, A., Roden, E.E., 2011. Iron in microbial metabolisms. Elements 7 inated aquifer. Geobiology 2 (1), 67–76.
(2), 89–93. Shi, L., Squier, T.C., Zachara, J., Fredrickson, J.K., 2007. Respiration of metal (hydr)oxides by
Koo, T., Jang, Y., Kogure, T., Kim, J.H., Park, B.C., Sunwoo, D., Kim, J., 2014. Structural and Shewanella and Geobacter: a key role for multihaem c-type cytochromes. Mol.
chemical modification of nontronite associated with microbial Fe(III) reduction: indi- Microbiol. 65 (1), 12–20.
cators of “illitization”. Chem. Geol. 377, 87–95. Sposito, G., 1984. The Surface Chemistry of Soils. Oxford University Press, New York.
Kostka, J.E., Stucki, J.W., Nealson, K.H., Wu, J., 1996. Reduction of structrural Fe(III) in Stucki, J.W., Kostka, J.E., 2006. Microbial reducion of iron in smectite. Compt. Rendus
smectite by a pure culture of Shewanella putrefaciens strain MR-1. Clay Clay Miner. Geosci. 338, 468–475.
44 (4), 522–529. Stucki, J.W., Goodman, B.A., Schwertmann, U., 1988. Iron in Soils and Clay Minerals. Reidel
Lalonde, K., Mucci, A., Ouellet, A., Gélinas, Y., 2012. Preservation of organic matter in sed- Publ. Co. NATO ASI Series C217.
iments promoted by iron. Nature 483, 198–200. Stucki, J.W., Lee, K., Goodman, B.A., Kostka, J.E., 2007. Effects of in situ biostimulation on
Li, Y., Vali, H., Sears, S.K., Yang, J., Deng, B., Zhang, C.L., 2004. Iron reduction and alteration iron mineral speciation in a sub-surface soil. Geochim. Cosmochim. Acta 71 (4),
of nontronite NAu-2 by a sulfate-reducing bacterium. Geochim. Cosmochim. Acta 68, 835–843.
3251–3260. Sun, Z., Zhou, H., Glasby, G.P., Yang, Q., Yin, X., Li, J., Chen, Z., 2012. Formation of Fe-Mn-Si
Liu, D., Dong, H., Bishop, M.E., Wang, H., Agrawal, A., Tritschler, S., Eberl, D.D., Xie, S., 2011. oxide and nontronite deposits in hydrothermal fields on the Valu Fa Ridge, Lau Basin.
Reduction of structural Fe(III) in nontronite by methanogen Methanosarcina barkeri. J. Asian Earth Sci. 43, 64–76.
Geochim. Cosmochim. Acta 75, 1057–1071. Taylor, K.G., Konhauser, K.O., 2011. Iron in earth surface systems: a major player in chem-
Liu, D., Dong, H., Bishop, M.E., Zhang, J., Wang, H., Xie, S., Wang, S., Huang, L., Eberl, D.D., ical and biological processes. Elements 7 (2), 83–88.
2012. Microbial reduction of structural iron in interstratidied illite-smectite minerals Vorhies, J.S., Gaines, R.R., 2009. Microbial dissolution of clay minerals as a source of iron
by a sulfate-reducing bacterium. Geobiology 10, 150–162. and silica in marine sediments. Nat. Geosci. 2, 221–225.
Liu, D., Dong, H., Zhao, L., Wang, H., 2014. Smectite reduction by Shewanella species as fa- Wang, C., Mao, N., Wu, F.T., 1980. Mechanical properties of clays at high pressure.
cilitated by cystine and cysteine. Geomicrobiol J. 31, 53–63. J. Geophys. Res. 85 (B3), 1462–1468.
Liu, D., Dong, H., Wang, H., Zhao, L., 2015. Low-temperature feldspar and illite formation Wang, F., Wang, J., Jian, H., Zhang, B., Li, S., Wang, F., Zeng, X., Gao, L., Bartlett, D.H., Yu, J.,
through bioreduction of Fe(III)-bearing smectite by an alkaliphilic bacterium. Chem. Hu, S., Xiao, X., 2008. Environmental adaptation: genomic analysis of the
Geol. 406, 25–33. piezotolerant and psychrotolerant deep-sea iron-reducing bacterium Shewanella
Liu, D., Dong, H., Agrawal, A., Singh, R., Zhang, J., Wang, H., 2016. Inhibitory effect of clay piezotolerans WP3. PLoS One 3, 1–12.
mineral on methanogenesis by Methanosarcina mazei and Methanothermobacter Weber, K., Achenbach, L.A., Coates, J.D., 2006. Microorganisms pumping iron: anaerobic
thermautotrophicus. Appl. Clay Sci. 126, 25–32. microbial iron oxidation and reduction. Nat. Rev. Microbiol. 4, 752–764.
Lovley, D.R., 1991. Dissimilatory Fe(III) and Mn(IV) reduction. Microbiol. Rev. 55, Williams, L.B., Hillier, S., 2014. Kaolins and health: from first grade to first aid. Elements 10
259–287. (3), 207–211.
Lovley, D.R., Coates, J.D., Blunt-Harris, E.L., Phillips, E.J.P., Woodward, J.C., 1996. Humic Wu, W., Wang, F., Li, J., Yang, X., Xiao, X., Pan, Y., 2013. Iron reduction and mineralization
substances as electron acceptors for microbial respiration. Nature 382, 445–447. of deep-sea iron reducing bacterium Shewanella piezotolerans WP3 at elevated hydro-
McCarty, D.K., Sakharov, B.A., Drits, V.A., 2009. New insights into smectite illitization: a static pressures. Geobiology 11 (6), 593–601.
zoned K-bentonite revisited. Am. Mineral. 94, 1653–1671. Xiao, X., Wang, P., Zeng, X., Bartlett, D., Wang, F., 2007. Shewanella psychrophila sp. nov.
Melton, E.D., Swanner, E.D., Behrens, S., Schmidt, C., Kappler, A., 2014. The interplay of and Shewanella piezotolerans sp. nov., isolated from west Pacific deep-sea sediment.
microbially mediated and abiotic reactions in the biogeochemical Fe cycle. Nat. Rev. Int. J. Syst. Evol. Microbiol. 57, 60–65.
Microbiol. 12, 797–808. Zhang, G., Kim, J., Dong, H., Sommer, A.J., 2007a. Microbial effects in promoting the smec-
Moore, D.M., Reynolds Jr., R.C., 1997. X-ray and the Identification and Analysis of Clay tite to illite reaction: role of organic matter intercalated in the interlayer. Am. Mineral.
Minerals. Oxford University Press, New York (Ch. 1, 2, & 3). 92, 1401–1410.
Morin, R., Silva, A.J., 1984. The effects of high pressure and high temperature or some Zhang, G., Dong, H., Kim, J., Eberl, D.D., 2007b. Microbial reduction of structural Fe3+ in
physical properties of ocean sediments. J. Geophys. Res. 89, 511–526. nontronite by a thermophilic bacterium and its role in promoting the smectite to illite
Murnane, R., Clague, D.A., 1983. Nontronite from a low-temperature hydrothermal sys- reaction. Am. Mineral. 92, 1411–1419.
tem on the Juan de Fuca Ridge. Earth Planet. Sci. Lett. 65, 343–352. Zhang, J., Dong, H., Liu, D., Fischer, T.B., Wang, S., Huang, L., 2012. Microbial reduction of
O'Reilly, S.E., Watkins, J., Furukawa, Y., 2005. Secondary mineral formation associated Fe(III) in illite-smectite minerals by methanogen Methanosarcina mazei. Chem.
with respiration of nontronite, NAu-1 by iron reducing bacteria. Geochem. Trans. 6, Geol. 292-293, 35–44.
67–76. Zhang, J., Dong, H., Liu, D., Agrawal, A., 2013. Microbial reduction of Fe(III) in smectite
Pevear, D.R., 1999. Illite and hydrocarbon exploration. Proc. Natl. Acad. Sci. 96, minerals by thermophilic methanogen Methanothermobacter thermautotrophicus.
3440–3446. Geochim. Cosmochim. Acta 106, 203–215.
Picard, A., Testermale, D., Wagenknecht, L., Hazael, R., Daniel, I., 2015. Iron reduction by
the deep-sea bacterium Shewanella profunda LT13a under subsurface pressure and
temperature conditions. Front. Microbiol. 5, 1–10.

You might also like