ArticlePDF Available

Nanoscale environments associated with bioweathering of a Mg-Fe-Pyroxene

Authors:

Abstract and Figures

Microorganisms are believed to create microenvironments leading to reaction products not predictable from equilibrium thermodynamics and to unique biomineral morphologies. Unambiguous evidence for such environments is, however, rare in natural samples. We have used scanning transmission x-ray microscopy and spectromicroscopy at the sub-40-nm scale, coupled with transmission electron microscopy, to examine bioweathering products on a meteoritic Fe-Mg-orthopyroxene colonized by a filamentous microorganism. Our measurements reveal an amorphous Al-rich layer beneath the microorganism, calcium carbonates of unique morphology intimately associated with polysaccharides adjacent to the microorganism, and regions surrounding the microorganism with different iron oxidation states. Our results confirm the presence of different microenvironments at this microorganism–mineral interface and provide unique nanometer-scale views of microbially controlled pyroxene weathering products. • aluminosilicate • biomineralization • geomicrobiology • STXM • Urey reaction
Content may be subject to copyright.
Nanoscale environments associated with
bioweathering of a Mg-Fe-pyroxene
Karim Benzerara
†‡
, Tae Hyun Yoon
, Nicolas Menguy
§
, Tolek Tyliszczak
, and Gordon E. Brown, Jr.
Surface and Aqueous Geochemistry Group, Department of Geological and Environmental Sciences, Stanford University, Stanford, CA 94305-2115;
§Laboratoire de Mine´ ralogie-Cristallographie, Unite´ Mixte de Recherche, 7590 Centre National de la Recherche Scientifique, and Institut de Physique du
Globe de Paris, 4 Place Jussieu, 75252 Paris Cedex, France; Chemical Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720; and
Stanford Synchrotron Radiation Laboratory, Stanford Linear Accelerator Center, 2575 Sand Hill Road, MS 99, Menlo Park, CA 94025
Communicated by W. G. Ernst, Stanford University, Stanford, CA, December 3, 2004 (received for review September 13, 2004)
Microorganisms are believed to create microenvironments leading
to reaction products not predictable from equilibrium thermody-
namics and to unique biomineral morphologies. Unambiguous
evidence for such environments is, however, rare in natural sam-
ples. We have used scanning transmission x-ray microscopy and
spectromicroscopy at the sub-40-nm scale, coupled with transmis-
sion electron microscopy, to examine bioweathering products on a
meteoritic Fe-Mg-orthopyroxene colonized by a filamentous mi-
croorganism. Our measurements reveal an amorphous Al-rich layer
beneath the microorganism, calcium carbonates of unique mor-
phology intimately associated with polysaccharides adjacent to the
microorganism, and regions surrounding the microorganism with
different iron oxidation states. Our results confirm the presence of
different microenvironments at this microorganism–mineral inter-
face and provide unique nanometer-scale views of microbially
controlled pyroxene weathering products.
aluminosilicate biomineralization geomicrobiology STXM
Urey reaction
Weathering of silicates at Earth’s surface has major impacts
on the environment, including the formation of soils and
the release of cationic nutrients (e.g., Fe and Mn) necessary for
the growth of microorganisms and plants (1). Other released
cations, such as Ca and Mg, can form Ca- andor Mg-carbonates,
resulting in sequestration of atmospheric CO
2
. This reaction,
known as the Urey reaction, has been suggested to control
Earth’s climate on geological time scales (2, 3). Pyroxenes are
among the most abundant silicates contributing to these reac-
tions (4). Abiotic dissolution of pyroxenes has been studied
extensively in the laboratory (e.g., see ref. 5), but significant
differences relative to dissolution rates measured in the field are
still not well understood (4, 6). The dissolution of Fe-containing
pyroxenes is even more complex because of the redox chemistry
of iron.
The role of microorganisms in the dissolution of silicates and
the cycling of Fe has been the focus of many experimental studies
(e.g., see ref. 7), some of which have shown that microorganisms
can radically modify the dissolution rates of silicates. They are
thought to do so by creating microenvironments in which pH and
other solution variables can dramatically differ from bulk con-
ditions (8, 9), by producing diverse metal-chelating organic
ligands (10), or by promoting metabolic redox reactions (11).
The pathways of microbially mediated Fe-silicate dissolution
reactions in nature remain, however, largely unexplored, partic-
ularly at the nanometer scale. Here we report new results on the
effects of a microorganism on the dissolution of an Fe
2
-Mg
orthopyroxene [(Mg
0.75
Fe
0.23
Ca
0.01
)(Si
0.99
Al
0.01
)O
3
] that was ex-
posed at Earth’s surface for 70 years. Although it may not be
broadly representative of weathering environments, this sample
from the Tatahouine meteorite, which fell in the southern
Tunisian desert in 1931, provides an unusual opportunity to
examine biotic reaction pathways in a well known time frame and
under defined conditions in an arid environment (12–15). Pre-
vious SEM and transmission electron microscopy (TEM) anal-
yses of the same sample (15) showed that some of the fragments
were colonized by filamentous microorganisms whose taxonomy
remains unknown. Evidence for a biological origin of this
filament comes from the carbon, nitrogen, and phosphorus
content as discussed in ref. 15. The microorganism is in contact
with the orthopyroxene and is bordered by a cluster of nanom-
eter-sized, rod-shaped calcite single crystals. Removal of the
calcite crystals showed that the underlying pyroxene surface is
pitted, suggesting some causal relationship involving the micro-
organism, calcite precipitation, and enhanced dissolution of the
orthopyroxene. This meteorite fragment was selected for further
analysis by using focused ion beam (FIB) milling to prepare an
ultrathin (80 nm) cross section through the microorganism, the
calcite crystals, and the pyroxene. Scanning transmission x-ray
microscopy (STXM) was used to perform high spatial and
energy resolution near-edge x-ray absorption fine structure
(NEXAFS) spectroscopy at the C K-edge, Al K-edge, and Fe
L
3
-edge at the microorganism–calcite–pyroxene interface.
These methods, in combination with TEM, provide remarkably
clear chemical-state-specific images of the same sample areas of
the microorganism–mineral interface at sub-40-nm spatial res-
olution and 0.1–0.3 eV (1 eV 1.602 10
19
J) energy
resolution that could not be obtained by a single method.
Materials and Methods
FIB Milling. FIB milling was performed with an FEI Model 200
TEM FIB system at the Universite´ Aix-Marseille III. The FIB
lift-out method was used to prepare the sample (see Fig. 4) as
described by Heaney et al. (16). The area studied by Benzerara
et al. (12) by SEM (figure 1a in ref. 12) could be easily located
by using the imaging capabilities of the FIB. A thin layer of
platinum was then deposited on the specimen across the filament
and the calcite cluster to protect them during the milling process.
The FIB system uses a Ga liquid metal ion source for milling. A
30 kV Ga
beam operating at 20 nA excavated pyroxene from
both sides of the Pt layer to a depth of 5
m. Before removal of
the thin slide, the sample was further thinned to 80 nm with a
glancing angle beam at much lower beam currents of 100 pA.
Finally, a line pattern was drawn with the ion beam along the side
and bottom edges of the thin section allowing its removal. The
15-
m5-
m80-nm slide was transferred at room pressure
with a micromanipulator onto the membrane of a C-coated 200
mesh copper grid.
TEM. TEM observations were carried out on a JEOL 2010F
microscope operating at 200 kV and equipped with a field
emission gun, a high-resolution Ultra High-Resolution pole
piece, and a Gatan GIF 200 energy filter.
Abbreviations: FIB, focused ion beam; NEXAFS, near-edge x-ray absorption fine structure;
STXM, scanning transmission x-ray microscopy; TEM, transmission electron microscopy.
To whom correspondence should be addressed. E-mail: benzerar@stanford.edu.
© 2005 by The National Academy of Sciences of the USA
www.pnas.orgcgidoi10.1073pnas.0409029102 PNAS
January 25, 2005
vol. 102
no. 4
979–982
GEOLOGY
STXM. STXM studies were performed at A LS branch line 11.0.2.2
with the synchrotron storage ring operating at 1.9 GeV and
200–400 mA stored current. Technical details about this new
branch line and its operation are provided in ref. 17, and recent
applications to colloids are presented in ref. 18. Energy calibra-
tion was made using the well resolved 3p Rydberg peak at 294.96
eV of gaseous CO
2
for C, the Fe L
3
maximum peak of the
Tatahouine pyroxene at 707.8 eV for Fe, and the first major
NEXAFS peak of montmorillonite at 1,567.7 eV. The detector
used in all measurements reported in this study was a photo-
multiplier tube with a phosphor scintillator. Stacks of images
were obtained by scanning in the xydirection at each energy
increment over the energy range of interest (280–320 eV for C,
1,555–1,580 eV for Al, and 705–730 eV for Fe). The contrast
results from differential absorption of x-rays, which is dependent
on the chemical composition of the sample. Normalization and
background correction (in particular the carbon signal from the
carbon film of the TEM grid) of the NEXAFS spectra were
performed by div iding each spectrum by a second spectrum from
a sample-free location on the grid. Spectral and spatial resolu-
tions were 0.1 eV and 40 nm, respectively, at the C K-edge and
Fe L
3
-edge, and 0.3 eV and 50 nm, respectively, at the Al K-edge.
AXIS2000 software (Version 2.1n, A. P. Hitchcock, McMaster Uni-
versity, Hamilton, ON, Canada) was used to align image stacks and
extract NEXAFS spectra from image stack measurements.
Results and Discussion
Calcium Immobilization in a Microbial Microenvironment. The C
K-edge NEXAFS spectrum of the filament is a complex mixture
of peaks (Fig. 1). Comparison with spectra from various biolog-
ical polymers (19) suggests the presence of polysaccharides (peak
at 288.6 eV) and proteins (peak at 288.2 eV) in the filament.
Together with the microbial-like morphology of this filament,
the presence of nitrogen and phosphorus (15) as well as bio-
chemical compounds, such as proteins, strongly support its
biological origin. The peak at 290.2 eV indicates the presence of
carbonates on or in the microorganism, suggesting fossilization.
C K-edge NEXAFS spectra of reference biological compounds
are scarce in the literature and are likely not representative of all
C-containing molecules present in this microorganism, which
helps explain why several spectral features could not be assigned.
Peaks at 284.3 eV, 286.4 eV, and 287.3 eV are thought to be
related to C functional groups present in either pristine microbial
molecules or molecules resulting from very early diagenesis. The
C K-edge NEXAFS spectrum of the calcite crystals is very
different from that of the microorganism, showing a major peak
at 290.2 eV (Fig. 2), which is characteristic of the
* resonance
of the CAO bond of carbonate groups. An additional peak at
288.6 eV was observed inside the calcite crystal cluster (Fig. 2)
and is indicative of carboxyl groups (
*CAO) likely associated
with polysaccharides (19, 20). Many authors have proposed the
involvement of organic molecules, either polysaccharides or
proteins, in the precipitation of calcium carbonates (e.g., see ref.
21). The nanometer-scale mixture of polysaccharides and car-
bonates observed on this sample suggests that calcite precipita-
tion occurred in a polysaccharide matrix likely released by the
filamentous microorganism. This suggestion is consistent with
several previous studies that presented models to explain how
polysaccharides could template carbonate precipitation (e.g., see
ref. 22). Moreover, the polysaccharides observed in this study
help explain the nanometer-sized, rod-shaped calcite crystals
found on the Tatahouine meteorite, which is an unusual mor-
phology for calcite crystals (12). Similar unusual morphologies
can be achieved when calcite growth occurs in the presence of
organic molecules, and polysaccharides in particular (23). More-
over, these calcite crystals are surrounded by an amorphous
calcium carbonate layer as shown by high-resolution TEM (12),
which is normally highly unstable under natural conditions but
can be stabilized in the presence of polysaccharides (24). Both
features thus provide potential biosignatures that could be useful
in the search for past life in earth and planetary materials.
Although some studies have suggested that abiotic dissolution of
Mg-silicates could be inhibited in the presence of carbonates
(e.g., see ref. 25), SEM and TEM observations on the Tata-
houine meteorite have shown that the filamentous microorgan-
ism enhanced the dissolution of the pyroxene below the calcite
Fig. 1. Spectromicroscopy analysis of filamentous microorganism. (A) TEM
image of the microorganism (arrows). Cc, calcite cluster; OPx, orthopyroxene.
The thin electron-dense layer covering the top of the cross-section is the
platinum layer deposited before FIB milling. (B) STXM image of the same area
at 270 eV. (C) C K-edge NEXAFS spectra from the microorganism, and refer-
ence calcite and organic molecules (from ref. 19); sodium alginate serves as a
model compound for polysaccharide, and albumin serves as a model com-
pound for proteins. Dashed lines at 288.2, 288.6, and 290.2 eV highlight the
most prevalent peaks of the model components.
Fig. 2. Spectromicroscopy analysis of the calcite nanocrystals. (A) TEM image
of the calcite nanocrystal cluster (see ref. 12). (B) Equivalent STXM image at
290.3 eV. (C) C K-edge NEXAFS spectra from the calcite cluster outlined on B,
and reference calcite and polysaccharide (sodium alginate) spectra (from ref.
19). Dashed lines at 288.6 and 290.2 eV highlight the mixture of calcite and
polysaccharides in the Tatahouine calcite cluster.
980
www.pnas.orgcgidoi10.1073pnas.0409029102 Benzerara et al.
cluster (12, 15). We suggest that this microorganism uniquely
coupled calcite precipitation and pyroxene dissolution by releas-
ing molecules that mediated Ca-carbonate precipitation. In
addition, we suggest that the same or other organic molecules
enhanced pyroxene dissolution. A similar Urey-like reaction
occurring in the presence of and promoted by microbes has been
indicated by several studies (26, 27). Such microbially mediated
reactions could be responsible for trapping some of the calcium
and magnesium released by pyroxene dissolution as well as
atmospheric CO
2
(2). The Urey reaction is classically considered
to be a two-step reaction on a time scale of hundreds of
thousands of years, with silicate weathering occurring on the
continents and carbonate precipitation occurring in the oceans
(2). Our results show that this reaction can take place on a very
small scale over a much shorter time scale.
Aluminum Fate in a Microorganism Microenvironment. The same
sample area was investigated by STXM imaging and NEXAFS
microspectroscopy at the Al K-edge (1,560 eV; Fig. 3). By taking
images below and above the Al K-edge, it was possible to obtain
an Al distribution map (Fig. 3). This map shows that Al is
preferentially concentrated beneath the microorganism. The Al
coordination number in this 100-nm-thick layer was determined
by acquiring an Al K-edge NEXAFS spectrum, which revealed
an absorption edge position at 1,564.5 eV and a broad peak
centered at 1,568 eV that are indicative of four- and sixfold-
coordinated Al (28). This result is consistent with a disordered
aluminosilicate, such as allophane (Fig. 3), and with TEM
observations showing that the Al-rich layer is amorphous at the
nanometer scale (15). Precipitation of clays at the surface of
pyroxenes has been reported by several workers (e.g., see ref.
29). Our results indicate that the local concentration of Al is
associated with the microorganism. One possible explanation for
this association is that the microorganism reduces the water
solid ratio, which has been used to explain local reprecipitation
of secondary products (30). This observation is also consistent
with the finding that microorganisms can mediate the precipi-
tation of aluminosilicates at their surface (31). Similar biologi-
cally induced microenvironments leading to localized etching
and aluminosilicate precipitation have been found in altered
oceanic volcanic glass by using optical microscopy (e.g., see ref.
32). In the present study, we observed the formation of an
Al-rich layer only beneath the microorganism, which may impact
the subsequent dissolution of the pyroxene. Indeed, such resid-
ual layers are potentially responsible for passivation of silicate
surfaces and inhibition of further silicate dissolution (4, 33). In
a Panglossian perspective, it is noted that the formation of this
layer beneath rather than surrounding the microorganism pro-
vides a stable anchor substrate without preventing the microor-
ganism from dissolving the neighboring substrate to take ad-
vantage of the release of essential nutrients from the pyroxene.
It has also been suggested (P. C. Bennett, personal communi-
cation) that the formation of this Al-rich layer may be beneficial
to the microorganism as it sequesters Al, which is a toxic metal
and interferes with iron-chelating siderophores. Characteriza-
tion of the trace element fluxes between this microorganism and
the pyroxene substrate would be of great help in evaluating these
suggestions.
Fe Oxidation at a Microorganism–Mineral Interface. Previous studies
of the abiotic dissolution of Fe(II)-bearing pyroxene have re-
vealed that Fe behavior is mostly controlled by redox conditions
(33). Under oxic conditions at near neutral pH, which is likely the
case for the Tatahouine pyroxene, Fe is oxidized, which may
passivate the surface of the pyroxene (34). In contrast, under
anoxic conditions, dissolution rates of Fe-pyroxenes are similar
to those of Fe-free pyroxenes (33). To verify these earlier
findings for dissolution of Fe-bearing pyroxenes under oxic
conditions, we studied the Tatahouine sample at the Fe L
3
-edge,
which allowed us to determine the iron oxidation state (e.g., see
ref. 35) in the microorganism–calcite–pyroxene microcosm at a
Fig. 4. Fe redox state analysis at the microbe–mineral interface. (A) TEM
image of the cross section showing the microorganism (arrow), the calcite
crystal cluster (Cc), and the orthopyroxene (Opx). (B) Equivalent STXM image
at 707.8 eV. (C) Iron L3-edge NEXAFS spectra from the pyroxene (area 1; see Fig.
2B), the calcite cluster (area 2), the microorganism (area 3), and reference
hematite, representing the Fe3endmember. Dashed lines represent the
positions of Fe L3maxima for Fe2and Fe3at 707.8 and 709.5 eV, respectively.
Fig. 3. Spectromicroscopy study of the microbe–mineral interface at the Al
K-edge. (A) STXM image of the microorganism (same area as the one observed
at the C edge; see Fig. 1) at 1,572 eV. (B) Corresponding TEM image. (C)
Aluminum elemental map obtained from the subtraction of the images of the
microorganism taken above and below the Al K-edge (at 1,572 and 1,560 eV,
respectively), showing high enrichment in Al only beneath the microorganism.
(D) Al K-edge NEXAFS spectrum from the Al-rich layer outlined in Fig. 2 A, and
reference spectrum of an allophane containing both four- and sixfold-
coordinated Al (from ref. 28).
Benzerara et al. PNAS
January 25, 2005
vol. 102
no. 4
981
GEOLOGY
spatial resolution of 40 nm (Fig. 4). Iron spectra taken on the
pyroxene (area 1) showed a major peak at 707.8 eV, indicative
of Fe
2
(Fig. 4). Fe-rich particles in the calcite cluster (area 2)
displayed a major peak at 709.5 eV, indicating that iron was
oxidized after its release by dissolution of the pyroxene (Fig. 4).
In the microorganism (area 3), however, the Fe L
3
-edge shows
a mixed iron valence (Fe
2
and Fe
3
) (Fig. 4). Spectra were
recorded several times over a several hour period, and all were
identical, indicating that beam damage was not responsible for
this observation in area 3. One possible explanation for variable
iron valence in and around the Tatahouine microorganism is that
the iron oxidation state was modified by this microorganism,
which creates a special microenvironment with a particular pO
2
andor modifies the pH, resulting in slower oxidation of iron.
Determining the taxonomy of the microorganism would help in
verifying this hypothesis, although such identification of a single,
partially fossilized microorganism, as in this case, would be
extremely difficult. Another possibility is that organic molecules
produced by the microorganism strongly bind dissolved Fe
2
and
thus inhibit its oxidation as suggested by several previous studies
(e.g., see ref. 36). Fe
3
is likely the stable form of iron at the
surface of the Tatahouine sand. Whatever mechanisms are
involved in iron redox behavior, the microorganism heavily
impacts iron oxidation dynamics in a microenvironment, result-
ing in a major modification of pyroxene reactivity, compared
with a purely abiotic environment. One consequence of the
stabilization of Fe
2
is the delay of entrapment of the microor-
ganism by Fe-oxides. Such stabilization could also provide a
low-energy-cost iron source.
Summary and Conclusions
A filamentous microorganism appears to have had a major
impact on the weathering of a meteoritic Fe
2
-Mg-orthopyrox-
ene during some period of its 70 years of exposure in the
southern Tunisian desert. Modifications of iron oxidation dy-
namics, calcium carbonate precipitation, and Al mobility by this
microorganism may have resulted in different dissolution rates
in regions surrounding the pyroxene separated by tens of nano-
meters. Our observations indicate that this microorganism cou-
pled carbonate precipitation and pyroxene weathering, provid-
ing a potential sequestration mechanism for CO
2
and producing
unique biosignatures in the form of rod-shaped nanocrystals of
calcite. The ability to characterize this microorganism– calcium
carbonate–pyroxene system at the sub-40-nanometer spatial
scale by using STXM and TEM methods provides a unique view
of how one type of microorganism controlled biomineral for-
mation (e.g., see ref. 37) as well as orthopyroxene weathering in
an arid environment. Although more systematic work on similar
samples from various environments will be needed to under-
stand the broader significance of the observations made here,
the methodology presented in this paper should be helpful in
assessing the importance of microorganisms in the evolution of
Earth’s surface chemistry and in identifying them in early Earth
and planetary materials.
We thank the CP2M members in Marseille who granted access to the
JEOL 2010F microscope and to the FEI Model 200 TEM FIB system,
A. P. Hitchcock for providing reference C K-edge spectra for albumin
and sodium alginate, Scott Fendorf (Stanford University) for a helpful
review of this manuscript, and Phil Bennett and two anonymous review-
ers for help in improving this article. The STXM studies were conducted
on branch line 11.0.2.2 at the Advanced Light Source, which is supported
by the Office of Science, Office of Basic Energy Sciences, Division of
Materials Sciences, and Division of Chemical Sciences, Geosciences, and
Biosciences of the U.S. Department of Energy at Lawrence Berkeley
National Laboratory under Contract DE-AC03-76SF00098. This work
was supported by National Science Foundation Grants CHE-0089215
(Stanford University Collaborative Research Activity in Environmental
Molecular Science on Chemical and Microbial Interactions at Environ-
mental Interfaces), CHE-0431425 (Stanford University Environmental
Molecular Science Institute), and EAR-9905755. K.B. thanks the French
Foreign Ministry for a Lavoisier Fellowship.
1. Marschner, H. (1995) The Mineral Nutrition of Higher Plants (Academic,
London).
2. Berner, R. A. (1995) Rev. Mineral. 31, 565–583.
3. Rothman D. H. (2001) Proc. Natl. Acad. Sci. USA 98, 4305– 4310.
4. Brantley, S. L. & Chen, Y. (1995) Rev. Mineral. 31, 119–172.
5. Oelkers, E. H. & Schott, J. (2001) Geochim. Cosmochim. Acta 65, 1219–
1231.
6. White, A. F. & Brantley, S. L. (2003) Chem. Geol. 202, 479–506.
7. Brantley, S. L., Liermann, L. J., Guynn, R. L., Anbar, A., Ic opini, G. A. &
Barling, J. (2004) Geochim. Cosmochim. Acta 68, 3189–3204.
8. Barker, W. W., Welch, S. A. & Banfield, J. F. (1997) Rev. Mineral. 35, 391– 428.
9. Little, B. J., Wagner, P. A. & Lewandowski, Z. (1995) Rev. Mineral. 35, 123–159.
10. Liermann, L. J., Kalinowski, B. E., Brantley, S. L. & Ferry, J. G. (2000)
Geochim. Cosmochim. Acta 64, 587–602.
11. Santelli, C. M., Welch, S. A., Westrich, H. R. & Banfield, J. F. (2001) Chem.
Geol. 180, 99–115.
12. Benzerara, K., Menguy, N., Guyot, F., Domin ici, C. & Gillet, P. (2003) Proc.
Natl. Acad. Sci. USA 100, 7438 –7442.
13. Barrat, J. A., Gillet, P., Lesourd, M., Blichert-Toft, J. & Poupeau, G. R. (1999)
Meteorit. Planet. Sci. 34, 91–97.
14. Barrat, J. A., Gillet, P., Lecuyer, C., Sheppard, S. M. F. & Lesourd, M. (1998)
Science 280, 412–414.
15. Benzerara K., Menguy, N., Guyot, F., Vanni, C. & Gillet, P. Geochim.
Cosmochim. Acta, in press.
16. Heaney, P. J., Vicenzi, E. P., Giannuzzi, L. A. & Livi, K. T. V. (2001) Am.
Mineral. 86, 1094–1099.
17. Tyliszczak, T., Warwick, T., Kilc oyne, A. L. D., Fakra, S., Shuh, D. K., Yoon,
T. H., Brown, G. E., Jr., Andrews, S., Chembrolu V., Strachan, J. & Acremann,
Y. (2004) AIP Conf. Proc. 705, 1356–1359.
18. Yoon, T. H., Johnson, S. B., Doyle, C. S., Benzerara, K., Tyliszczak, T., Shuh,
D. K. & Brown, G. E., Jr. (2004) Langmuir 20, 10361–10366.
19. Lawrence, J. R., Swerhone, G. D. W., Leppard, G. G., Araki, T., Zhang, X.,
West, M. M. & Hitchcock, A. P. (2003) Appl. Environ. Microbiol. 69, 5543–5554.
20. Myneni, S. C. B. (2002) Rev. Mineral. Geochem. 49, 485–579.
21. Kirkland, B. L., Lynch, F. L., Rahnis, M. A., Folk, R. L., Molineux, I. J. &
McLean, R. J. C. (1999) Geology 27, 347–350.
22. Arp, G., Reimer, A. & Reitner, J. (2003) J. Sediment. Res. 73, 105–127.
23. Weiner, S. & Addadi, L. (1997) J. Mater. Chem. 7, 689 –702.
24. Aizenberg, J., Lambert, G., Weiner, S. & Addadi, L. (2002) J. Am. Chem. Soc.
124, 32–39.
25. Wogelius, R. A. & Walther, J. V. (1991) Geochim. Cosmochim. Acta 55,
943–954.
26. Ferris, F. G., Wiese, R. G. & Fyfe, W. S. (1994) Geomicrobiol . J. 12, 1–13.
27. Roberts, J. A., Bennett, P. C., Gonzalez, L. A., Macpherson, G. L. & Milliken,
K. L. (2004) Geology 32, 277–280.
28. Ildefonse, P., Kirkpatrick, R. J., Montez, B., Calas, G., Flank, A. M. & Lagarde,
P. (1994) Clays Clay Miner. 42, 112–121.
29. Banfield, J. F., Barker, W. W., Welch, S. A. & Taunton, A. (1999) Proc. Natl.
Acad. Sci. USA 96, 3404 –3411.
30. Casey, W. H., Westrich, H. R., Banfield, J. F., Ferruzzi, G. & Arnold, G. W.
(1993) Nature 366, 252–256.
31. Ueshima, M. & Tazaki, K. (2001) Clays Clay Miner. 49, 292–299.
32. Fisk, M. R., Giovannoni, S. J. & Thorseth, I. H. (1998) Science 281, 978 –980.
33. Schott, J. & Berner, R. A. (1983) Geochim. Cosmochim. Acta 47, 2233–2240.
34. Siever, R. & Woodford, N. (1979) Geochim. Cosmochim. Acta 43, 717–724.
35. Peng, G., Van Elp, J., Jang, H., Que, L., Armstrong, W. H. & Cramer, S. P.
(1995) J. Am. Chem. Soc. 117, 2515–2519.
36. Neubauer, S. C., Emerson, D. & Megonigal, J. P. (2002) Appl. Environ.
Microbiol. 68, 3988–3995.
37. Chan, C. S., De Stasio, G., Welch, S. A., Girasole, M., Frazer, B. H., Nesterova,
M. V., Fakra, S. & Banfield, J. F. (2004) Science 303, 1656–1658.
982
www.pnas.orgcgidoi10.1073pnas.0409029102 Benzerara et al.
... Overall, it is possible to preserve cell-mineral interfaces at least in some cases. For example, early on, Benzerara, Menguy, Guyot, Vanni, and Gillet (2005), Benzerara, Yoon, Menguy, Tyliszczak, and Brown (2005) produced an electron-transparent lamella cut across a fungus cell grown over a silicate grain. This lamella was then investigated by high-resolution TEM and STXM, evidencing the formation and dissolution of minerals triggered by this cell and the onset of a chemical microenvironment. ...
... In the latter case, the surface right underneath the fungal biofilm remained mostly crystalline and was strongly etched and weathered, indicating enhanced olivine dissolution. , Benzerara, Yoon, et al., 2005 correlated TEM imaging, SAED, EDXS and EELS analyses with synchrotron-based scanning transmission X-ray microscopy (STXM) to measure speciation of Fe, C, Ca, C and O by XANES spectroscopy in an environmental assemblage of fungi grown on pyroxene crystals. They also suggested enhanced mineral dissolution together with calcium carbonate precipitation associated with the extracellular polymers produced by the microorganisms. ...
Chapter
Biofilms can interact with minerals, by either dissolving the substrates on which they are attached and/or inducing mineral precipitation within their extracellular matrix. Diverse scientific fields have studied such mineral-biofilm interactions as they have numerous and broad implications, including the aging and conservation of monument stones, the formation of caries, a control on soil fertility, the geological CO2 cycle and water quality. Despite the importance of mineral-biofilms interactions, they remain difficult to investigate since this task requires biological, crystallographic and chemical information at high spatial resolution and at the same time over large areas in order to grasp heterogeneities within the biofilms. The hardness of minerals associated with the softness and hydrated nature of cells further complicate the use of standard methods. Here we review multiple spatially resolved techniques that have been applied to biofilm-mineral assemblages or may be particularly useful for this purpose, including light (OCT and CLSM) and electron (SEM, TEM, FIB) microscopies, vibrational and X-ray spectroscopies, as well as atomic force microscopy and vertical scanning interferometry. We also emphasize the promising developments that may push such investigations to a new era.
... At the Fe L-edge, the references Fe II _ref and Fe III _ref present four main peaks at 708 eV (Fe L3 edge), 709.8 eV (Fe L3 edge), near 720-721 eV (Fe L2 edge) and near 723 eV (Fe L2 edge) (Fig. 1a) [37,38]. The peaks intensities of these two references are different which, in principle, makes it possible to quantify the Fe II -Fe III ratio as was done on silicates systems [42]. ...
... a Reference of iron(II) chloride (Fe II _ref ) and iron(III) chloride (Fe III _ref ) at the Fe L-edge; b gelatin (G) at the N K-edge and c C K-edge of gallic acid reference (Ac), iron-gallate ink precipitate (IGI), gelatin (G) and raw linen rag fiber (fiber). Attribution was based on several previous studies [34,[37][38][39][40][41] Gelatin is the only compound of the system that contains nitrogen. Hence, N K-edge mapping will be specific of the presence of gelatin. ...
Article
Full-text available
Iron gall Inks corrosion causes paper degradation (browning, embrittlement) and treatments were developed to tackle this issue. They often include resizing with gelatin to reinforce the paper and its cellulosic fibers (of diameter approx. 10 µm). This work aimed at measuring the distribution of ink components at the scale of individual paper fibers so as to give a better understanding of the impact of gelatin (re-)sizing on iron gall ink corrosion. For this purpose, scanning transmission X-ray microscopy (STXM) was used at the Canadian light source synchrotron (CLS, Saskatoon). This technique combines nano-scale mapping (resolution of 30 nm) and near edge X-ray absorption fine structure (NEXAFS) analysis. Fe L-edge measurements enabled to map iron distribution and to locate iron(II) and iron(III) rich areas. N K-edge measurement made it possible to map gelatin distribution. C K-edge measurements allowed mapping and discrimination of cellulose, gallic acid, iron gall ink precipitate and gelatin. Three fibers were studied: an inked fiber with no size, a sized fiber that was afterwards inked and an inked fiber sprayed with gelatin. Analysis of gelatin and ink ingredients distribution indicated a lower amount of iron inside the treated cellulosic fiber, which may explain the beneficial effect of gelatin on iron gall ink corrosion.
... In the case of lava caves, the basalt host rock contains abundant mineral substrates with reduced iron (e.g., olivine, pyroxene) that has been shown to provide an energy source for chemolithotrophic microorganisms [38,39]. Microbial "mining" for reduced iron and manganese in the carbonate matrix has also been implicated in wall corrosion in limestone caves [20,40] and in other iron-manganese rich minerals (e.g., [41]). Further, biological iron Apatite was the most common secondary mineral observed. ...
... In the case of lava caves, the basalt host rock contains abundant mineral substrates with reduced iron (e.g., olivine, pyroxene) that has been shown to provide an energy source for chemolithotrophic microorganisms [38,39]. Microbial "mining" for reduced iron and manganese in the carbonate matrix has also been implicated in wall corrosion in limestone caves [20,40] and in other iron-manganese rich minerals (e.g., [41]). Further, biological iron reduction has been proposed to play a role in the formation of iron ore caves [42]. ...
Article
Full-text available
Biovermiculations are uniquely patterned organic rich sediment formations found on the walls of caves and other subterranean environments. These distinctive worm-like features are the combined result of physical and biological processes. The diverse microbial communities that inhabit biovermiculations may corrode the host rock, form secondary minerals, and produce biofilms that stabilize the sediment matrix, thus altering cave surfaces and contributing to the formation of these wall deposits. In this study, we incubated basalt, limestone, and monzonite rock billets in biovermiculation mixed natural community enrichments for 468–604 days, and used scanning electron microscopy (SEM) to assess surface textures and biofilms that developed over the course of the experiment. We observed alteration of rock billet surfaces associated with biofilms and microbial filaments, particularly etch pits and other corrosion features in olivine and other silicates, calcite dissolution textures, and the formation of secondary minerals including phosphates, clays, and iron oxides. We identified twelve distinct biofilm morphotypes that varied based on rock type and the drying method used in sample preparation. These corrosion features and microbial structures inform potential biological mechanisms for the alteration of cave walls, and provide insight into possible small-scale macroscopically visible biosignatures that could augment the utility of biovermiculations and similarly patterned deposits for astrobiology and life detection applications.
... STXM has been used to investigate the rhizosphere in few recent studies, for studying OC distribution within bacterial biofilms at fine scale [129][130][131] soil microaggregates [12,26,51] and bacteria-soil mineral interfaces [132]. Along with STXM, NEXAFS has been used to study oxidative damage of lignocellulosic materials [133] in bleaching and if such technique could be used in soil, it may help in quantifying the degradation of plant cells [134]. ...
Chapter
Full-text available
Soil is a highly heterogenous system where a number of physical, chemical and biological processes are taking place. The study of these processes requires analytical techniques. The electromagnetic radiations in the form spectroscopy, X-Ray diffraction, magnetic resonance etc. have been used in the field of soil analysis since decades. The study of soil nutrients, mineralogy, organic matter and complex compounds in soils use these techniques and are successful tools till date. But these come with a limitation of lesser spatial and spectral resolution, time consuming sample preparation and destructive methods of study which are mostly ex-situ. In contrast to the conventional spectroscopic techniques, the synchrotron facility is of high precision and enables non-destructive study of the samples to a nano scale. The technique uses the high intensity synchrotron radiation which is produced in a special facility, where the electrons are ejected using very high voltage and accelerated in changing magnetic field, at a speed of light resulting in a very bright radiation that enables a very précised study of the subject. For example, in studying the dynamics of P and N in soils, SR aided XAS are used to study the K-edge spectra of these nutrients , without any matrix interference, which used to be a problem in conventional SEM, IR or NMR spectroscopy. These radiations provide high energy in GeV, which imparts high sensitivity and nanoscale detection. Basically, the SR facility improves the precision of the existing spectroscopic techniques. This chapter discusses how the Synchrotron radiations aid to improve precision in various field of soil analysis such as, carbon chemistry, nutrient dynamics, heavy metal and contaminant speciation and rhizosphere study. However, the technique also come with major limitations of requirement of very high skill for preparation of samples, inadequate availability of references for studies related to absorption spectrum and control of radiation damage. Applications and limitations of the technique thoroughly reviewed in this chapter with an aim to provide a brief idea of this new dimension of soil analysis.
... Echoing the paradigm proposed by Hazen et al. (2008), the coevolution of the geo-and biospheres may favor microbial communities spe-cifically adapted to each mineral surface, possibly harboring microorganisms with efficient weathering ability (Uroz et al., 2015). Microenvironments are generated at the silicate-microbe interface (Benzerara et al., 2005), where locally aggressive conditions can remarkably increase silicate weathering rates (Bonneville et al., 2009;Li et al., 2016), as reported in numerous laboratory experiments involving microbial cultures (Uroz et al., 2015). However, quantitative upscaling of laboratory results to natural settings is not straightforward. ...
Article
Full-text available
Chemical weathering produces solutes that control groundwater chemistry and supply ecosystems with essential nutrients. Although microbial activity influences silicate weathering rates and associated nutrient fluxes, its relative contribution to silicate weathering in natural settings remains largely unknown. We provide the first quantitative estimates of in situ silicate weathering rates that account for microbially induced dissolution and identify microbial actors associated with weathering. Nanoscale topography measurements showed that fungi colonizing olivine [(Mg,Fe)2SiO4] samples in a Mg-deficient forest soil accounted for up to 16% of the weathering flux after 9 mo of incubation. A local increase in olivine weathering rate was measured and attributed to fungal hyphae of Verticillium sp. Altogether, this approach provides quantitative parameters of bioweathering (i.e., rates and actors) and opens new avenues to improve elemental budgets in natural settings.
... The dissolution of enstatite is non-stoichiometric under acidic conditions (pH = 2) and its dissolution is faster than at higher pH. Few studies were dedicated to enstatite or pyroxene bioweathering (Berner and Schott 1982;Herrera et al. 2004;Benzerara et al. 2004Benzerara et al. , 2005Bassez 2017). In those studies, the main biological effect was to slow down the alteration kinetics presumably by protection mechanisms of the mineral surface. ...
Article
Understanding the transformations of highly reduced enstatite chondrites (EC) in terrestrial environments, even on very short timescales, is important to make the best use of the cosmochemical and mineralogical information carried by these extraterrestrial rocks. Analogs of EC meteorites were synthesized at high pressure and high temperature. Then, their aqueous alterations, either abiotic or in the presence of the bacteria Acidithiobacillus ferrooxidans or Acidithiobacillus thiooxidans, were studied under air, at pH ~2, 20 °C, and atmospheric pressure. They stayed in shaken batch reactors for 15 days. Reference experiments were carried out separately by altering only one mineral phase among those composing the synthetic EC (i.e., sulfides: troilite or Mg‐Ca‐rich sulfides, enstatite, and Fe70Si30). Composition of the alteration aqueous media and microstructures of the weathered solids were monitored by inductively coupled plasma atomic emission spectroscopy and by scanning electron microscopy, respectively. Alteration sequence of the different mineral components of the synthetic EC was found to occur in the following order: magnesium‐calcium sulfides > troilite > iron‐silicon metallic phase > enstatite regardless of the presence or absence of the microorganisms. Such small biological effects might be due to the fact that the alteration conditions are far from biologically optimal, which is likely the case in most natural environments. The exposed surfaces of an EC meteorite falling on Earth in a wet and acidic environment could lose within a few hours their Ca‐ and Mg‐rich sulfides (oldhamite and niningerite). Then, in <1 week, troilite and kamacite could be altered. In a wet and acidic environment, only the enstatite would remain intact and would weather on a much slower geological timescale.
Article
Full-text available
Purpose Root-derived low-molecular-weight organic acids (LMWOAs) can impact the decomposition of soil organic matter (SOM) after being released into soils. However, the influence of individual LMWOA (e.g., oxalic acid) inputs on the destabilization of physicochemically protected SOM remains largely unknown. Methods Using artificial roots in a firmly controlled rhizosphere system, we daily added oxalic acid solutions to soils collected from two subalpine coniferous forests (a 70‐year‐old spruce plantation and a 200‐year‐old spruce‐fir dominated forest) and incubated the soils for over 25 days. Results The addition of oxalic acid significantly decreased the concentrations of iron bound in metal–organic complexes (Fe-MOCs), aluminum bound in metal–organic complexes (Al-MOCs), iron bound in short-range order phases (Fe-SROs), and iron bound in short-range order phases (Al-SROs) by 35%, 13%, 16%, and 30%, respectively, across the two forest soils. This result indicated that the oxalic acid addition promoted the destabilization of physicochemical-protected SOM. This destabilization of protected SOM was mainly caused by breaking crosslinking between carboxylic groups and multivalent cations and the release of aromatic carbon (C) from mineral-organic associations, as indicated by the concurrently decreased zeta potential and the prominently featured resonances assigned to aromatic functional groups in the corresponding spectra of the near edge X-ray absorption fine structure after the addition of oxalic acid. In addition, compared to that of the spruce plantation, the addition of oxalic acid induced greater changes in the metal pools (Fe and Al) bound in MOCs and SROs in the spruce-fir forest, which indicated that the oxalic acid-induced destabilization of physicochemical-protected SOM might also be regulated by native soil properties. Conclusions Our study demonstrates that the input of LMWOAs to soils could stimulate the destabilization of physicochemical-protected SOM, which is presumably involved in the disruption of mineral-organic associations by breaking the crosslinking bonds and releasing aromatic C. The destabilization of physicochemically protected SOM may accelerate SOM decomposition, and thus, the input of LMWOAs to soils has important ecological implications for the biogeochemical cycle in terrestrial ecosystems.
Chapter
This article overviews connections between nanoscale weathering and geomorphology. Nanoscale processes are on one side of a fundamental threshold between the coarser microscale (micrometers and up) and the finer nanoscale with different molecular dynamics. Nanoscale processes impact a variety of geomorphic research including Arctic and alpine mineral decay, biotic weathering as an explanation for deviations from Goldich's weathering series, carbon sequestration related to silicate dissolution, case hardening, detachment limited erosion, dirt cracking, geochemical pollution, the meteoric ¹⁰Be cosmogenic nuclide, rock coating behavior, silt production, and tafoni.
Article
Full-text available
Nontronite and microbes were detected in the surface layers of deep-sea sediments from Iheya Basin, Okinawa Trough, Japan. Nontronite, an Fe-rich smectite mineral, was embedded in acidic polysaccharides that were exuded by microbial cells and electron microscopy showed that the nontronite layers were apparently oriented in the polysaccharide materials. We propose that the formation of nontronite was induced by the accumulation of Si and Fe ions from the ambient seawater and that extracellular polymeric substances (EPS) served as a template for layer-silicate synthesis. Experimental evidence for this hypothesis was obtained by mixing a solution of polysaccharides (dextrin and pectin) with ferrosiliceous groundwater. After stirring the mixture in a sealed vessel for two days, and centrifuging, Fe-rich layer silicates were identified within the precipitate of both the dextrin and pectin aggregates, whereas rod-shaped or spheroidal Si-bearing iron hydroxides were found in the external solution. Microbial polysaccharides would appear to have affected layer-silicate formation.
Article
Full-text available
Direct light and electron microscopic studies show that cyanobacterial cells serve as nucleation sites for carbonate mineral precipitation in a variety of fresh to saline‐alkaline lakes on the Cariboo Plateau in central British Columbia, Canada, and in mineralized crusts on weathered basalt in Iceland. The carbonate minerals found in association with the cyanobacteria were extremely fine‐grained, and invariably occurred on the external surfaces of the cells. Carbonate mineralogy was variable, ranging from calcite to magnesite, depending on differences in lake and groundwater chemistry (i.e., saturation state of the water with respect to individual carbonate minerals). In microcosm experiments, phototrophic cyanobacterial growth increased alkalinity and the degree of oversaturation with respect to calcite. Calculated values for the saturation state of calcite and magnesite in Cariboo Plateau natural waters exhibited two distinct trends, with (1) high magnesite saturation values in areas where the weathering of magnesium olivine‐rich basalt bedrock determines water chemistry, and (2) high calcite saturation values where bedrock is a mix of basic lava flows, limestone, argillite, and chert. Similar calculations for Iceland show that cold surface waters are generally oversaturated with respect to calcite, as expected for the weathering of calcium plagioclase‐rich lava. These observations of microbial carbonate precipitation in the Caribou Plateau region of British Columbia, Canada, and on the Budarhaun lava plain in Iceland suggest that weathering of silicate minerals in bedrock is biogeochemically coupled to the deposition of carbonate minerals by microorganisms. This process may provide a sink for atmospheric carbon dioxide in terrestrial environments.
Article
Argon ion milling is the conventional means by which mineral sections are thinned to electron transparency for transmission electron microscope (TEM) analysis, but this technique exhibits significant shortcomings. In particular, selective thinning and imaging of submicrometer inclusions during sample milling are highly problematic. We have achieved successful results using the focused ion beam (FIB) lift-out technique, which utilizes a 30 kV Ga+ ion beam to extract electron transparent specimens with nanometer scale precision. Using this procedure, we have prepared a number of Earth materials representing a range of structures and compositions for TEM analysis. We believe that FIB milling will create major new opportunities in the field of Earth and planetary materials microanalysis, particularly with respect to ultraprecious mineral and rock samples.
Article
Acidithiobacillus ferrooxidans are commonly present in acid mine drainage (AMD). A. ferrooxidans derive metabolic energy from oxidation of Fe²⁺ present in natural acid solutions and also may be able to utilize Fe²⁺ released by dissolution of silicate minerals during acid neutralization reactions. Natural and synthetic fayalites were reacted in solutions with initial pH values of 2.0, 3.0 and 4.0 in the presence of A. ferrooxidans and in abiotic solutions in order to determine whether these chemolithotrophic bacteria can be sustained by acid-promoted fayalite dissolution and to measure the impact of their metabolism on acid neutralization rates. The production of almost the maximum Fe³⁺ from the available Fe in solution in microbial experiments (compared to no production of Fe³⁺ in abiotic controls) confirms A. ferrooxidans metabolism. Furthermore, cell division was detected and the total cell numbers increased over the duration of experiments. Thus, over the pH range 2–4, fayalite dissolution can sustain growth of A. ferrooxidans. However, ferric iron released by A. ferrooxidans metabolism dramatically inhibited dissolution rates by 50–98% compared to the abiotic controls.
Article
The dissolution rates of Fo100 and Fo91 in aqueous solutions in the pH range 2–12.4 at 25°C have been measured using fluidized bed and batch reactors. Rates depend upon the pH, the partial pressure of CO2, and the presence of organic ligands. At low PCO2 (≤10−4.5 atm) with no organic ligands present, the rate of olivine dissolution, R, is given by , where aH+ is the activity of H+ in solution. However in basic solutions, when the partial pressure of CO2 is equal to atmospheric levels (PCO2 = 10−3.5 atm), olivine dissolution rates are nearly pH independent throughout the pH range 6–12 and are about equal to the minimum rate of dissolution under CO2 purged conditions. At pH 11 the presence of atmospheric levels of CO2 reduces the dissolution rate by over an order of magnitude (to 10−14.1 mol cm−2 s−1). Apparently, positive charge on the olivine surface can be neutralized by increasing PCO2. In contrast, experiments conducted in the acidic and near neutral pH ranges indicate that organic ligands chelate surface Mg causing an increase in the olivine dissolution rate when present. Organic ligand effects are greatest in the near neutral pH domain. For example, at pH 4 dissolution rates are increased by 0.75 log units (to 10−12.25 mol cm−2 s−1) in solutions of 10−3 molar ascorbic acid or 0.05 molar potassium acid phthalate over rates measured in organic free solutions. The chelation effect becomes less important as pH decreases. Rates at pH 2 in the presence of these organic ligands are indistinguishable from those measured without organics.
Article
Organic molecules are found everywhere and play an important role in almost all biogeochemical processes occurring on the surface of the Earth (Aiken et al. 1985; Thurman 1985; Schwartzenbach et al. 1993; Senesi and Miano 1994). They are found in soluble and insoluble phases, coatings on mineral and colloidal particles, and in gas phase molecules in soils, sediments and aquatic systems. The activities of macro- and micro-fauna and flora release organic molecules of various sizes and composition. Photochemical reactions in the atmosphere also add certain small chain molecules to the organic carbon content in the environment. Significant compositional variations occur in natural organic molecules, which include small chain carboxylic acids, alcohols and amino acids; and polymeric, polyfunctional and polydisperse macromolecules such as humic and fulvic acids. The behavior of small chain molecules and their influence on different geochemical reactions is well understood. However, understanding of the chemistry of biopolymers and their role in different biogeochemical processes in the environment is poor, which may be attributed to the unavailability of instrumentation to examine the chemistry of natural organic molecules in their pristine state. Two important properties that dictate the behavior of natural organic molecules and biopolymers in the environment are: functional group chemistry and (macro)molecular structure (Schnitzer 1991). Evaluation of these two properties is complicated by the compositional and structural heterogeneity of the naturally occurring organic molecules, and their ability to form intramolecular and intermolecular H-bonds, which further modify their structure and chemical reactivity. These two properties are interrelated and one influences the other. The chemical composition of natural waters (pH, ionic composition and concentration, redox conditions) and soil and sediment particle surface chemistry (composition, coordination environment, number of reactive groups) also modify their behavior. It is not well understood how each of these environmental variables influences the …
Article
The crater lake of the small volcanic island Satonda, Indonesia, is unique for its red-algal microbial reefs thriving in marine-derived water of increased alkalinity. The lake is a potential analogue for ancient oceans sustaining microbialites under open-marine conditions. Current reef surfaces are dominated by living red algae covered by non-calcified biofilms with scattered cyanobacteria and diatoms. Minor CaCO3 precipitates are restricted to the seasonally flooded reef tops, which develop biofilms up to 500 mum thick dominated by the cyanobacteria Pleurocapsa, Calothrix, Phormidium, and Hyella. Microcrystalline aragonite patches form within the biofilm mucilage, and fibrous aragonite cements grow in exopolymer-poor spaces such as the inside of dead, lysed green algal cells, and reef framework voids. Cementation of lysed hadromerid sponge resting bodies results in the formation of "Wetheredella-like" structures. Hydrochemistry data and model calculations indicate that CO2 degassing after seasonal mixis can shift the carbonate equilibrium to cause CaCO3 precipitation. Increased concentrations of dissolved inorganic carbon limit the ability of autotrophic biofilm microorganisms to shift the carbonate equilibrium. Therefore, photosynthesis-induced cyanobacterial calcification does not occur. Instead, passive, diffusion-controlled EPS-mediated permineralization of biofilm mucus at contact with the considerably supersaturated open lake water takes place. In contrast to extreme soda lakes, the release of Ca2+ from aerobic degradation of extracellular polymeric substances does not support CaCO3 precipitation in Satonda because the simultaneously released CO2 is insufficiently buffered. Subfossil reef parts comprise green algal tufts encrusted by micro-stromatolites with layers of fibrous aragonite and an amorphous, unidentified Mg-Si phase. The microstromatolites probably formed when Lake Satonda evolved from seawater to Ca2+-depleted raised-alkalinity conditions because of sulfate reduction in bottom sediments and pronounced seasonality with deep mixing events and strong CO2 degassing. The latter effect caused rapid growth of fibrous aragonite, while Mg-Si layers replaced the initially Mg-calcite-impregnated biofilms. This could be explained by dissolution of siliceous diatoms and sponge spicules at high pH, followed by Mg-calcite dissolution and Mg-silica precipitation at low pH due to heterotrophic activity within the entombed biofilms.
Article
Organisms have been producing mineralized skeletons for the past 550 million years. They have evolved many different strategies for improving these materials at almost all hierarchical levels from Ångstroms to millimetres. Key components of biological materials are the macromolecules, which are intimately involved in controlling nucleation, growth, shaping and adapting mechanical properties of the mineral phase to function. One interesting tendency that we have noted is that organisms have developed several strategies to produce materials that have more isotropic properties. Much can still be learned from studying the principles of structure–function relations of biological materials. Some of this information may also provide new ideas for improved design of synthetic materials.
Article
Iron L-edge X-ray absorption and X-ray magnetic circular dichroism (XMCD) spectroscopy have been used to study the electronic structure of dinuclear iron-oxo complexes with different types of magnetic and electronic interactions between the iron sites. Trapped-valence systems exhibit L-edges with clear multiplet structure. The L-edges of trapped-valence Fe[sup II]Fe[sup III] complexes such as [Fe[sup III,II]2(salmp)[sub 2]][sup [minus]] and [Fe[sub 2][sup III,II](bpmp)([mu]-O[sub 2]CC[sub 2]H[sub 5])[sub 2]][sup 2[center dot]] can be interpreted as the sum of distinct Fe(II) and Fe(III) component spectra. Furthermore, an atomic multiplet theory including adjustable ligand field splittings can successfully simulate the Fe(II) and Fe(III) X-ray absorption. Reasonable ligand field parameters are obtained by optimizing the correspondence between calculated and experimental spectra. The XMCD for the [Fe[sub 2][sup III,II](bpmp)([mu]-O[sub 2]CC[sub 2]H[sub 5])[sub 2]][sup 2] complex is also reported; it exhibits an interesting magnetic field dependence that reflects the weak magnetic coupling between Fe(II) and Fe(III) ions. In contrast with the trapped-valence complex spectra, the L-edge spectrum for the electronically delocalized complex, [Fe[sub 2](Me[sub 3]tacn)[sub 2]-([mu]-OH)[sub 3]](BPh[sub 4])[sub 2]-2MeOH, exhibits a broad L-edge spectrum with poorly resolved multiplet structure. 35 refs., 4 figs.