ArticlePDF AvailableLiterature Review

The Impact of Host Rock Geochemistry on Bacterial Community Structure in Oligotrophic Cave Environments.

Authors:

Abstract and Figures

Despite extremely starved conditions, caves contain surprisingly diverse microbial communities. Our research is geared toward understanding what ecosystems drivers are responsible for this high diversity. To asses the effect of rock fabric and mineralogy, we carried out a comparative geomicrobiology study within Carlsbad Cavern, New Mexico, USA. Samples were collected from two different geologic locations within the cave: WF1 in the Massive Member of the Capitan Formation and sF88 in the calcareous siltstones of the Yates Formation. We examined the organic content at each location using liquid chromatography mass spectroscopy and analyzed microbial community structure using molecular phylogenetic analyses. In order to assess whether microbial activity was leading to changes in the bedrock at each location, the samples were also examined by petrology, X-ray diffraction (XRD) and scanning electron microscopy with energy dispersive X-ray spectroscopy (SEM-EDX). Our results suggest that on the chemically complex Yates Formation (sF88), the microbial community was significantly more diverse than on the limestone surfaces of the Capitan (WF1), despite a higher total number of cells on the latter. Further, the broader diversity of bacterial species at sF88 reflected a larger range of potential metabolic capabilities, presumably due to opportunities to use ions within the rock as nutrients and for chemolithotrophic energy production. The use of these ions at sF88 is supported by the formation of a corrosion residue, presumably through microbial scavenging activities. Our results suggest that rock fabric and mineralogy may be an important driver of ecosystem function and should be carefully reviewed when carrying out microbial community analysis in cave environments.
Content may be subject to copyright.
Available online at www.ijs.speleo.it
International Journal of Speleology
Ofcial Journal of Union Internationale de Spéléologie
The Impact of Host Rock Geochemistry on Bacterial Community Structure
in Oligotrophic Cave Environments.
Hazel A. Barton1, Nicholas M. Taylor1, Michael P. Kreate2, Austin C. Springer2,
Stuart A. Oehrle3 and Janet L. Bertog2.
INTRODUCTION
Caves, with limited exception, form through the
erosional processes of water. By the time caves are
enlarged sufciently to allow human access, the water
has (generally) departed, leaving the cave exposed to
an oxygenated atmosphere (Klimchouk et al., 2000).
Without sunlight energy, the entry of nutrients into the
system becomes a function of the geology and depth
of the cave; signicant organic input is limited to the
entrance zone and areas fed by surface water entering
the system through faults and fractures (Klimchouk
et al., 2000). Due to extremely low biomass in these
environments and the difculty in extracting DNA
1.Department of Biological Sciences, Northern Kentucky
University, Highland Heights, KY 41099, United States of
America.
2.Department of Physics & Geology, Northern Kentucky
University, Highland Heights, KY 41099, United States of
America.
3.Department of Chemistry, Northern Kentucky University,
Highland Heights, KY 41099, United States of America.
Barton H. A., Taylor N. M., Kreate M. P., Springer A. C., Oehrle S. A. and Bertog J. L. 2007. The impact of host rock geochemistry on
bacterial community structure in oligotrophic cave environments. International Journal of Speleology, 36 (2), 93-104. Bologna (Italy).
ISSN 0392-6672.
Despite extremely starved conditions, caves contain surprisingly diverse microbial communities. Our research is geared toward
understanding what ecosystems drivers are responsible for this high diversity. To asses the effect of rock fabric and mineralogy,
we carried out a comparative geomicrobiology study within Carlsbad Cavern, New Mexico, USA. Samples were collected from
two different geologic locations within the cave: WF1 in the Massive Member of the Capitan Formation and sF88 in the calcareous
siltstones of the Yates Formation. We examined the organic content at each location using liquid chromatography mass spectroscopy
and analyzed microbial community structure using molecular phylogenetic analyses. In order to assess whether microbial activity
was leading to changes in the bedrock at each location, the samples were also examined by petrology, X-ray diffraction (XRD) and
scanning electron microscopy with energy dispersive X-ray spectroscopy (SEM-EDX). Our results suggest that on the chemically
complex Yates Formation (sF88), the microbial community was signicantly more diverse than on the limestone surfaces of the
Capitan (WF1), despite a higher total number of cells on the latter. Further, the broader diversity of bacterial species at sF88
reected a larger range of potential metabolic capabilities, presumably due to opportunities to use ions within the rock as nutrients
and for chemolithotrophic energy production. The use of these ions at sF88 is supported by the formation of a corrosion residue,
presumably through microbial scavenging activities. Our results suggest that rock fabric and mineralogy may be an important driver
of ecosystem function and should be carefully reviewed when carrying out microbial community analysis in cave environments.
Keywords: caves, geomicrobiology, geochemistry, phylogenetics, oligotrophy
Abstract:
Received 26 April 2007; Revised 14 May 2007; Accepted 22 May 2007
from chemically complex geologic samples, most
studies of microbial activity in cave environments
have tended to examine areas of measurable energy
input (Angert et al., 1998; Barton & Luiszer, 2005;
Bottrell et al., 1991; Culver, 1982; Groth et al.,
2001; Sarbu et al., 1996). In a recent study we used
molecular phylogenetics to determine what, if any,
microbial activity was occurring within an oligotrophic
cave without measurable energy input (Barton et al.,
2004). Our results suggested that a diverse microbial
ora subsisted in this oligotrophic environment, in
contrast to previous cultivation studies from similar
environments (Groth & Saiz-Jimenez, 1999; Groth et
al., 1999). Further, the microbial community within
this environment appeared to subsist by using barely
perceptible carbon and energy sources; these included
organics entering the system through percolation,
or the presence of volatile organic molecules within
the atmosphere (Barton et al., 2004). The presence
of bacterial phylotypes with identity to organisms
capable of carrying out iron oxidation suggested
the use of reduced iron as an energy source, while
International Journal of Speleology 36 (2) 93-104 Bologna (Italy) July 2007
94
a high proportion of nitrogen assimilating organisms
suggested a source for nitrogen (Laiz et al., 1999). A
broad phylogenetic distribution of bacterial species
has been identied by other investigators in similarly
oligotrophic environments (Chelius & Moore, 2004;
Osman et al., 2005).
Together, our data led us to hypothesize that
the large diversity of microorganisms found
in oligotrophic cave environments may reect
mutualistic interactions to support community
growth under such starved conditions (Barton &
Jurado, 2007); due to the complex nature of the
organic carbon and inorganic energy sources,
not one organism is capable of carrying out all
the energetically favorable reactions necessary to
support growth (Juttner, 1984; Laiz et al., 1999).
Rather, energetic restrictions allow certain reactions
to proceed only through a close interaction with
species that remove intermediates, allowing energy
conservation in what would otherwise be an
endothermic reaction (Schink, 2002). Mutualism has
been described for microbial communities carrying
out anaerobic ethanol fermentation, methane
oxidation and the breakdown of complex aromatic
compounds (Schink, 2002). Such interactions may
also be a central issue in the unculturability of most
microorganisms in the environment; mutualistic
interactions make many organisms recalcitrant to
cultivation, where appropriate growth conditions
may be dependent on specic interactions with
other species (Grotenhuis et al., 1991; Mohn &
Tiedje, 1992).
To further examine drivers of microbial diversity
in starved environments, this study attempts to
determine what role rock fabric and mineralogy
plays in community diversity under oligotrophic
conditions. This was done by carrying out a
comparative analysis of two microbial communities
within Carlsbad Cavern, New Mexico, USA, where
bacterial species exist on disparate geologic surfaces.
We used a molecular and geochemical approach
to examine community structure at each location
and compared microbial interactions with the rock
matrix of the cave.
MATERIALS AND METHODS
Sample sites and geology
Carlsbad Cavern was formed in the Capitan Reef complex
by hypogenic sulfuric acid speleogenesis, with a postulated
biogenic origin (Engel et al., 2004; Hill, 1990; Palmer, 2000).
The Carlsbad cave system is mostly within the Capitan
limestone, where the relatively impermeable iron-rich, silty
Yates Formation traps oxygenated groundwater and releases
it into the Capitan Formation (Palmer, 2000). Upper portions
of the cave are also located in the Yates Formation, which
contains numerous calcareous siltstones, sandstones and
secondary minerals. The cave is located is in a desert area,
does not contain any surface streams and is not prone to
ooding.
Samples for collection were identied within Carlsbad
Caverns based on a number of parameters, including
geologic location, altered bedrock or secondary
mineralization (Fig. 1). Therst site, WF1, is along the
Main Corridor and located in the limestone of the Massive
Member of the Capitan Formation (CaCO3), with an average
annual temperature of 12.5˚C (Forbes 2000) and a relative
humidity (RH) of 95%, measured using an RH300 Digital
Psychrometer (Extech Instruments, Waltham, MA). The
second site, sF88, is located within the Yates Formation,
with the sample collection site directly above a calcareous
siltstone bed (Fig. 1), with an average annual temperature of
16.3˚C and measured RH of 99% (Forbes 2000). The Yates is
comprised a ne-grained, laminated pisolitic dolomite in thin
beds, inter-layered with thin layers of calcareous red quartz
siltstones andne-grained sandstones (Borer & Harris,
1991; Brown & Loucks, 1993; DuChene, 2000; Mutti &
Simo, 1993). The Yates is rich in magnesium and iron, with
its red color due to the presence of hematite (Fe2O3), which
is not generally detectable by scanning electron microscopy
energy dispersive X-ray spectroscopy (SEM-EDX) (Borer &
Harris, 1991). Three 5 g rock samples were collected from
each location using a sterilized Dremel drill tool and each
sample was preserved in an appropriate manner for the
subsequent tests: DNA extraction in 70% alcohol / -20°C;
chemical samples were collected in gamma-irradiated clean
tubes and stored at 4°C; rock samples were collected in 50
ml plastic tubes.
Fig. 1. Prole (line plot facing north) of the Carlsbad Cavern cave system (approximately 48 km of passage is represented) with the corresponding
geologic units of the Capitan Reef complex overlain. The two sample locations (WF1 and sF88) are indicated by the lled circles. The entrance and
‘Big Room’ are designated. Courtesy of the Cave Resource Ofce, Carlsbad Caverns National Park.
Hazel A. Barton, Nicholas M. Taylor, Michael P. Kreate, Austin C. Springer, Stuart A. Oehrle and Janet L. Bertog.
International Journal of Speleology, 36 (2), 93-104. Bologna (Italy). July 2007
95
Chemical Analysis
Total organic carbon (TOC) was measured by
extracting crushed rock with dH2O and then
determining the TOC g-1 of rock material using a
Shimadzu TOC-VCSN analyzer at Waters Laboratory,
Western Kentucky University, KY. Analyses of sample
extracts for organic carbon were carried out using
high performance liquid chromatography-mass
spectrometry (HPLC/MS). The system consisted of an
Alliance 2695 HPLC system, 2996 photodiode array
detector and a ZQ single quadrapole mass spectrometer
(all equipment was from Waters Corp., Milford, MA).
Approximately 100mg of sample from each of the sites
was extracted in 1 ml of a 50/50 water and acetonitrile
solution by sonication (all solvents were of HPLC
grade or better). Samples were than allowed to stand
and settle prior to the top (clear) layer extracted and
analyzed by HPLC/MS. A standard gradient was run
using an Xterra MS C18 column (2.1X100mm 3.5µm
particle) using a formic acid and acetonitrile gradient
over 30 minutes.
DNA extraction
DNA extractions were carried out in a laminar-
ow hood, using aseptic techniques and aerosol
resistant tips to reduce the chance of contamination
from outside sources (Barton et al., 2006). Unless
stated otherwise, all chemicals were obtained from
Sigma-Aldrich (St. Louis, MO) and reagents used
were prepared from Fluka ultrapure DNase/RNase
Free water, followed by ltration through a 0.2 µm
cellulose lter to prevent contamination. In cases
where contaminating DNA may be introduced from
reagents, these were subjected to 3000 µJ cm-1 of
UV radiation using a Stratalinker 2400 (Stratagene,
La Jolla, CA). To extract the small amount of DNA
present in the rock, a modied bead beating method
was used.
To extract the DNA, approximately 0.5g of sample
was crushed using a ame-sterilized plattner’s mortar
and pestle (Humboldt Manufacturing, Norridge, IL).
To this 500µl 2X buffer EA [200 mM Tris (pH 8.0),
300 mM EGTA, 200 mM NaCl], 3 mg/ml lysozyme
and 10 µg/ml poly-dIdC were added, and incubated
at 37˚C for 30 min. Proteinase K (to 1.2 mg/ml) and
sodium dodecyl sulfate (SDS to 0.3% wt/vol) were
added, mixed gently and incubated at 50°C for 30
min. Subsequently, 200µl of 20% SDS and 500µl
phenol-chloroform-isoamyl alcohol (24:24:1) was
added before disruption using a Mini-bead beater
(Biospec, Bartlesville, OK) on low setting for 2 min
and high for 30 s. Samples were centrifuged at 13,000
x g in a micro-centrifuge for 3 min at 4°C to deposit
the sample debris; the supernatant (approximately
700-800µl) was then removed and the DNA by the
addition of 2 µg poly-dIdC, 0.3 M sodium acetate and
2 volumes of cold ethanol. Isolated DNA was further
puried by dialysis against 100 ml of 20 mM EGTA
at 4°C for 4 hours in a Silde-A-Lyzer mini dialysis
unit (3500 MWCO; Pierce, Rockford, IL) to remove
any remaining calcium carbonate. The concentration
of the nal DNA preparation was determined using
a Nanodrop ND-1000 spectrophotometer (Nanodrop
Technologies, Wilmington, DE).
Polymerase Chain Reaction (PCR) and cloning
To isolate individual 16S rRNA gene clones, PCR
amplication was used). The bacterial 16S rRNA gene
specic 27F (5’ AGA GTT TGA TCC TGG CTC AG
– 3’) and universal 805R (5’ – GAC TAC CAG GGT ATC
TAA T 3’) primers were used in reaction mixtures
containing 1 X PCR buffer (Perkin Elmer), 2.5 mM
MgCl2, 200 μM of each deoxynucleoside triphosphate,
300 nM of each forward and reverse primer, and
0.025 U of AmpliTaq Gold (Perkin Elmer) per μl.
Reaction mixtures were incubated on a Mastercycle
Gradient thermal cycler (Eppendorf Scientic) at 94˚C
for 12 min for initial denaturation and activation of
the AmpliTaq Gold. PCR was then carried out with 34
cycles of 94ºC 30 s, 58ºC 30 s, 70ºC 1 min 30 s, and
a nal extension period of 70ºC 2 min. PCR products
were quantied by electrophoresis using a 1.2% wt/
vol agarose gel containing ethidium bromide, puried
using a Qiagen PCR clean up kit (Qiagen, Valencia,
CA) and cloned into an pCR2.1-TOPO cloning vector
according to the manufacturers recommendations
(Invitrogen, Carlsbad, CA).
Screening of rDNA clones by restriction length
polymorphism (RFLP) and DNA sequencing
The 16S rRNA gene inserts were PCR re-amplied
using 100 ng T3 forward and T7 reverse primers
under standard conditions and amplied using 94ºC
for 4 min for initial denaturation, then 38 cycles of
94ºC 1 min, 52ºC 45 s, 72ºC 1 min, with an extension
period of 72ºC for 8 min. PCR products were then
digested using HindPI1 and MspI restriction enzymes
in NEB buffer 2 (New England Biolabs, Beverly, MA).
The restriction digest was incubated at 37ºC for 2
hours before being run on a 2% wt/vol SeaKem LE
agarose gel (FMC BioProducts) and visualized with
ethidium bromide staining with UV illuminescence.
The unique RFLP patterns were grouped visually
and a representative was selected for sequencing.
Sequencing was carried out using the Thermo
Sequenase Cycle Sequencing kit (USB, Cleveland,
Ohio) according to the manufacturer’s guidelines. For
areas that were problematic due to regions of high
GC content, a SequiTherm Excell II DNA sequencing
kit (Epicenter Technologies, Madison, WI) was used.
Sequencing was carried out using uorescently
labeled sequencing primers M13 and T7 on a Long
ReadIR 4200 DNA sequencer (Li-Cor, Lincoln, NE),
which achieved approximately an 800 base rRNA gene
insert in both the forward and reverse directions.
Phylogenetic Analysis
Sequences were compared to available databases
by use of the BLAST (Basic Local Alignment Search
Tool) network service [http://www.ncbi.nlm.nih.gov/
BLAST; Altschul et al., 1997] Partial sequences of the
16S rRNA gene were compiled using the AlignIR 2.0
Fragment Assembly and Contig Editor software (Li-Cor,
Inc). Compiled sequences were examined for chimeric
The Impact of Host Rock Geochemistry on Bacterial Community Structure in Oligotrophic Cave Environments.
International Journal of Speleology, 36(2), 93-104. Bologna (Italy). July 2007
96
sequences by use of the CHIMERA_CHECK program
[http://rdp.cme.msu.edu/html/analyses.html]
and by phylogenetic branching order discrepancies.
Before further phylogenetic analysis, those sequences
displaying similar BLAST hits were directly compared
using the pairwise BLAST alignment tool [http://
www.ncbi.nlm.nih.gov/blast/bl2seq/bl2.html]. Any
sequences that demonstrated 98% identity toward
each other were considered representatives of the same
phylotype and grouped accordingly; the remaining
sequences screened for contaminants against
the LBC database (Barton et al., 2006). Sequence
alignments were carried out using the ARB Software
Package [http://mpi-bremen.de/molecol/arb], with
additional sequences from the Ribosomal Database
Project (Maidak et al., 2000). Sequence alignments
used for phylogenetic inference were minimized by
use of the Lane Mask, which removes hypervariable
regions of the 16S rRNA gene sequence from the
analysis for Bacterial data sets. Due to the size of the
16S rRNA clones isolated in this study (~800 bp), only
representative sequences from position 40 to 790 (E.
coli numbering) were used in subsequent phylogenetic
analyses. All presented dendrograms were constructed
by use of ARB with evolutionary distance (neighbor-
joining) and parsimonious (heuristic) algorithms.
The robustness of inferred topologies was tested by
bootstrap resampling (1000 replicates) of phylogenetic-
trees, calculated for both algorithms using PAUP*
software (Sinauer Associates Inc., Sunderland, MA).
The sequences obtained from the 16S rRNA gene
clones in this study were deposited in the Genbank
database, accession numbers DQ066600-DQ066618
and DQ228711-DQ228720.
Statistical Approaches
The statistical analyses were performed using
EstimateS version 7.5.0 (Colwell, 2005). Each clone
represented a separate sample without replacement,
and 100 randomizations were performed to obtain the
Chao2 estimator for each sample size (Chao et al.,
2005; Gotelli & Colwell, 2001; Hughes et al., 2001).
Using the singletons and doubletons calculated for
each sample collection by EstimateS, we used the
log transformation of Chao to calculate the 95%
condence intervals (Chao, 1987; Chao et al., 2005).
Rarefaction curves were plotted using SigmaPlot for
Windows Version 7.0 (SPSS Inc., Point Richmond,
CA).
Scanning Electron Microscopy (SEM)
Samples were xed prior to analysis in 4%
paraformaldehyde/PBS. Biological samples were
washed in 70% ethanol, and dehydrated in an ethanol
series to 100%. Samples were dried in a critical point
dryer using liquid CO2 before examination under
environmental SEM conditions using a FEI Quanta
200 ESEM with a Princeton Gamma Tech Avalon
Microanalysis system and environmental secondary as
well as backscatter electron detectors. Bulk chemistry
of the host rock was determined by crushing samples
and examining them by SEM-EDX using and EDAX
brand EDX mounted on a Philips XL30 TMP scanning
electron microscope.
Geologic analysis
Whole rock samples were separated into an outer
residue layer and interior bedrock, powdered using a
Spex Certiprep 8515 shatterbox and analyzed using
a Rigaku Ultima III XRD under air-dried conditions.
Samples were run from 2 (theta) to 70 (theta) with
a step size of 0.05 (theta) and a count time of 2
seconds. For petrographic analysis, whole rock
samples were sectioned, embedded in resin and cut
and ground to 0.03mm for thin section analysis by
Vancouver Petrographic Ltd. Samples were analyzed
using a Nikon E400Pol polarizing microscope.
RESULTS
We began by identifying two distinct areas in Carlsbad
Cavern that appeared to have limited organic carbon
input, but quite different geologic settings (Fig. 1).
WF1 is considered a chemically simple environment
predominately calcium carbonate in chemistry, while
sF88 is chemically complex, reecting shoal and
eolian deposition process that interlayered pisolitic
dolomite with crystalline volcanic and metamorphic
rock fragments. As a result, the Yates contains various
accessory minerals, including hematite, magnetite
(Fe3O4), tourmaline [(Ca,Na)(Li,Mg,Al)(Al,Fe,Mn)6(BO
3)3(Si6O18)(OH)4], zircon (ZrSiO4), rutile (TiO2), apatite
[Ca5(PO4)3(F,Cl,OH)] and epidote [Ca2(Al,Fe)3(SiO4)3(OH
)]. To conrm these chemistries, we carried out EDX
analysis of the host rock (Table 1), which conrmed
the limestone nature of WF1 and the presence of
quartz siltstone (SiO), magnesium, aluminum,
potassium and iron at sF88. The other elements
previously identied in the Yates (Li, Mn, B, etc.) were
below the minimal detectable limits of this technique.
At WF1 (Capitan), the rock on which microbial species
are growing does not appear to have undergone any
signicant rock fabric changes, while at sF88 (Yates)
varied colored corrosion residues were seen on the
surface. We counted the number of microbial cells at
each location by uorescent microscopy (Barton et
al., 2006). At WF1, there were 1.75 x 106 cells g-1 of
wall rock, while at sF88 we observed 6.48 x 105 cells
g-1 of material, which match the cell numbers seen
in similarly oligotrophic cave environments (Barton et
al., 2006). No eukaryotic predators were observed by
microscopy from either location, ruling out predation
as a driver of diversity (Hahn & Hoe, 2001).
Organic Chemistry and LC/MS analysis
Given the different lithology and mineralogy at
each site, for example the Yates in known to be rich
in organics and is a major reservoir for oil and gas
in the Delaware Basin (Borer & Harris, 1991), we
carried out a qualitative analysis to determine if
there was a difference in the type of organic material
present at each location. This would also rule out the
presence of organics as a result of surface spills from
the commercial facilities (sewage tanks, fuel tanks,
etc.) above the cave. Analysis was done of each of
International Journal of Speleology, 36 (2), 93-104. Bologna (Italy). July 2007
Hazel A. Barton, Nicholas M. Taylor, Michael P. Kreate, Austin C. Springer, Stuart A. Oehrle and Janet L. Bertog.
97
Fig. 2. Consensus dendogram of the 16S rRNA gene sequence phylotypes identied from the Carlsbad clone libraries WF1 (white boxes) and sF88
(black boxes). Phylogenetic analyses were carried out using both distance (Neighbor-Joining) and parsimonious (Heuristic) searches, with the
robustness of inferred topologies determined by bootstrap analysis (1000 replicates); the consensus dendogram of both methods is shown. Branch
points supported (bootstrap values >70%) in both phylogenetic analyses are indicated by closed circles, while marginal branch support (bootstrap
values >50% but >70%) in both analyses are show by an open circle. The bar indicates 10% sequence divergence.
The Impact of Host Rock Geochemistry on Bacterial Community Structure in Oligotrophic Cave Environments.
International Journal of Speleology, 36(2), 93-104. Bologna (Italy). July 2007
98
the sites using liquid chromatography coupled mass
spectrometry (LC/MS). A preliminary analysis of
the LC/MS peaks indicates that the organic matter
present is low (measured as 3 µg g-1 at sF88), although
slightly higher at WF1 (results not shown). The data
suggest that the organic material is of a phenolic and
aromatic nature, which is in accord with the structure
of the organic material commonly found in soils and
the postulated origin of much of the organic carbon
observed in caves (Saiz-Jimenez & Hermosin, 1999;
Sylvia et al., 1999).
Molecular Phylogenetic Analysis of WF1 versus
sF88 Communities
In order to determine whether the differences
in geologic chemistry affected the structure of the
microbial communities found at WF1 and sF88 we
carried out a molecular phylogenetic study. Due to the
difculty of obtaining DNA from calcium rich samples,
along with the small biomass associated with these
extremely starved environments, we have developed
a new extraction protocol for calcium rich samples
from caves (Barton et al., 2006). While this protocol
allows us to reproducibly obtain DNA from extremely
low-biomass environments, we still routinely obtain
less than 500 ng of community DNA/g of material. We
therefore use the 8F and 805R 16S rRNA gene primer
set for amplication, which provides the most reliable
PCR amplication at these low DNA concentrations.
While the subsequent rDNA product is short (~800 bp
in length), it still provides sufcient information for
statistically signicant phylogenetic placement (Nei et
al., 1998).
Clone libraries were created for both sF88 and WF1
locations by ligating the PCR product into the plasmid
vector and transforming into chemically competent E.
coli cells; for sF88 144 clones were isolated, while at
WF1, 96 representative clones were used. The 16S
rRNA gene sequence of each clone library was screened
by RFLP analysis to identify unique phylotypes. The
nal clone libraries contained 49 unique phylotypes
for sF88 and 38 unique phylotypes for WF1 and were
groups into operational taxanomic units (OTUs),
demonstrating >98% identity for tree building. The
sequences of these phylotypes were compared with
the NCBI database and the closest cultivated relative
was identied (Table 2).
Surprisingly, many of the phylotypes we identied
shared a greater degree of identity with previously
cultivated species than in a past cave study (Barton et
al., 2004); however, this may simply reect the increase
in size of the 16S rRNA gene sequence database. In order
to conrm the identity of these identied phylotypes, they
were phylogenetically aligned using the ARB sequence
analysis program, followed by statistical analysis of the
resultant dendogram using the PAUP* software program.
The consensus tree for each location conrmed the
phylogenetic placement of the identied species (Fig. 2).
Interestingly, the distribution of phylotypes identied
correlates well with previous cave environments and
similarly starved locations (Barton et al., 2004; Chelius
& Moore, 2004; Osman et al., 2005). It is also interesting
to note that many of the sF88 phylotypes share the
closest identity to 16S rRNA gene sequences identied
in the WF1 library, suggesting the shared ancestry of
the two communities within this cave environment.
While the identity of unique phylotypes at sF88 and
WF1 demonstrates a similar distribution among the
bacterial divisions, the dendogram does not reect the
relative abundance of the phylotypes identied in each
location, which are represented in Fig. 3. It is interesting
that this comparative pie-chart demonstrates a much
greater species distribution within the chemically
complex sF88 environment, even while there is a higher
absolute number of bacterial cells at the WF1 site. In
order to determine whether there were statistically
signicant differences in community structure between
each site, we created rarefaction curves using the Chao
2 non-parametic estimator to determine true species
richness (Fig. 4). These rarefaction curves did indeed
suggest that there were differences in the absolute
diversity between the two microbial communities;
however, the 95% condence levels suggest that the
sample sizes need to be increased to determine the
signicance of these differences.
Geologic Samples: Thin-sections and X-ray powder
diffractrometry (XRD)
One of the most striking differences between WF1
and sF88 sites was the presence of a corrosion residue
Fig.3. Pie-charts representing species distribution at each location
within Carlsbad Cavern. The distribution of all divisions are
represented by each cloned phylotype from the WF1 (88 clones) and
sF88 (144 clones) libraries.
Fig.4. Rarefaction curves of observed operational taxonomic units
(OTUs) represented by individual phylotypes identied within this
study. Each rarefaction curve was calculated from the variance of the
number of OTUs drawn in 100 randomizations at each sample size.
International Journal of Speleology, 36 (2), 93-104. Bologna (Italy). July 2007
Hazel A. Barton, Nicholas M. Taylor, Michael P. Kreate, Austin C. Springer, Stuart A. Oehrle and Janet L. Bertog.
99
at the more starved sF88 site. The two localities were
compared for mineralogic alteration using thin-section
petrography, XRD and SEM-EDX. Petrologically, both
localities were quite different, reecting the difference in
composition between the Capitan and Yates formations
(Fig. 5A and 5B). SEM images conrmed a calcite
bedrock with an apparent microbial biolm at WF1,
but no signicant corrosion surface (Fig. 5A and 5C).
At sF88, however, the surface of the rock underwent a
number of mineralogical and crystallographic changes,
resulting in a poorly consolidated corrosion residue
(Fig. 5B and 5D). This corrosion residue is comprised
of dolomite recrystallized into coarser crystals, with
ne-grained clay minerals and other opaque minerals
present between them.
Microbial species often change the chemical nature
of the environment on which they live through
catabolic processes (Baneld & Nealson, 1997). In
order to determine if such transformations were
occurring at WF1 or sF88 we carried out an SEM-
EDX analysis of insoluble particulate matter in the
rock. In order to identify such minerals, we extracted
the rock in each location with 1M hydrochloric acid
to remove the overwhelming carbonate minerals and
examined individual particulate grains. At WF1 this
material is comprised of clay particles (representative
EDX spectra in Fig. 5E and 5F) normally associated
with the Massive Member of the Capitan Formation
while at sF88 this material comprised of iron oxides
and elemental iron (representative EDX spectra in Fig.
5G and 5H). These iron forms are too small and/or
amorphous to be detected in our XRD analyses (Fig.
6), suggesting a biogenic origin. These results support
the theory that iron oxidation may be one mechanism
of energy production at the sF88 location, indicating
that the microorganisms living in these environments
may be acquiring energy from the host rock itself.
In order to compare the changes in geologic structure
at the two localities, samples were analyzed using
comparative XRD (Fig. 6). At each location, the sample
collected was broken down into two components; the
interior bedrock and the surface layers. The interior
rock was >1 cm away from any observable surface
feature (as demonstrated by the thin-section analysis)
Fig.5. Petrographic thin-sections, SEM and EDX analysis of rock samples from WF1 and sF88. Samples were collected and subjected to thin
sectioning and petrographic analysis at WF1 (A) and sF88 (B). The interior bedrock is mineralogically similar to the outer layer at WF1 (A), while
the corrosion residue lls the entire eld-of-view of the sF88 sample (B). SEM analysis reveals a similar surface mineralogy at WF1 (C), where
individual calcite crystals covered with a biolm material are seen on the surface. A ne, powdery residue predominates on the surface at sF88 (D).
EDX analyses of acid extracts at each location demonstrate the predominance of clay particles at WF1 (E and F) and iron oxides and elemental
iron at sF88 (G and H)
The Impact of Host Rock Geochemistry on Bacterial Community Structure in Oligotrophic Cave Environments.
International Journal of Speleology, 36(2), 93-104. Bologna (Italy). July 2007
100
and was considered representing the bedrock
mineralogy. Each layer was clearly marked in hand
sample, allowing for segregation of the layers prior to
powdering for XRD. The results (Fig. 6) demonstrated
that, in agreement with our thin-section analysis,
that there were no signicant changes in the
mineral structure of the surface rock at WF1, when
compared with the host-rock matrix; however, there
was an accumulation of clay sized particles. We have
seen such clays accumulating at other sites within
Carlsbad Cavern that demonstrate biogenic activity
(Bertog et al., unpublished results). Heating of the
sample to 350˚C for 30 min led to the loss of the clay
peak, suggesting kaolinite. At sF88 there appeared
to be more signicant mineral changes; dolomite in
the bedrock had been removed with an increase in
the relative abundance of the non-soluble bedrock
material, such as quartz and other silicates (Fig. 6)
as would be expected if the calcareous cements of
the Yates Formation had been dissolved. SEM images
of sF88 conrmed a crystallographic change in the
corrosion residue, with a ne (<1µm) powdery residue
(Fig. 5D).
DISCUSSION
The majority of caves contain little available carbon,
making them an ideal environment in which to study
oligotrophic microbial interactions and geochemical
processes on exposed surfaces (Laiz et al., 1999). Such
geomicrobial activity is thought to be indicated by the
presence of corrosion residues: areas of fabric and
mineralogical change in the bedrock, characterized
by a color change and softening or powdering of the
rock (Boston et al., 2001; Canaveras et al., 2001;
Northup et al., 2003). While these residues do contain
an observable microbial population, a mechanism of
formation remains to be determined (Canaveras et al.,
2001; Northup et al., 2003).
At the geochemically simple WF1 site (Capitan), the
microbial community is dominated by members of
the Actinobacteria, a broad class of high G+C, gram-
positive bacteria found predominantly within soil. Of
these, representative phylotypes of the Pseudonocardia
appear to be the most abundant, representing half of
all the identied Actinobacteria and over 80% of the
total community of bacteria found at this location
(Fig. 3). While the identity between the predominant
phylotype and the next closest cultivated species is
only 97%, we can postulate on a general function of
this species in the environment (Achenbach & Coates,
2000; Pace, 1997). Members of the Pseudonocardia
are aerobes that demonstrate a wide metabolic range
for the degradation of complex plant matter, such
as cellulose, suggesting that the community at WF1
is primarily using soil detritus for growth (Dworkin,
2002). Interestingly, other phylotypes identied at
WF1 share similarity to Acinetobacter johnsonii, able
to mobilize phosphate from inorganic sources, and
Comamonas spp., which degrade a number of nitrogen-
containing aromatic compounds, with the release of
usable nitrate and ammonia (Dworkin, 2002; Itoh &
Shiba, 2004). Both of these groups similarly display
saprophytic lifestyles and are routinely found in
the environment under nutrient limiting conditions
(Dworkin, 2002).
In contrast to the relatively simple microbial
diversity identied at WF1, the clone library generated
at the more geochemically complex sF88 site (Yates)
was more diverse, with representatives from the
Alpha-, Beta- and Gammaproteobacteria (Fig. 3);
very similar in structure to other oligotrophic cave
environments (Barton et al., 2004; Chelius & Moore,
2004). Among the phylotypes identied, there was
signicant representation by members of the genera
Brevundimonas, Massilia and Stenotrophomonas;
26%, 18% and 17% respectively. Representative
Brevundimonas spp. are from the Caulobacter family,
which are oligotrophic organisms able to adapt to
extremely starved environments (Dworkin, 2002; Li
et al., 2004). Members of the genus Massilia are able
to utilize a large number of carbohydrates and other
complex organic molecules as carbon and energy
sources, with the subsequent production of acids
(Dworkin, 2002). Such activity may explain some of
the signicant structural changes observed in the
Yates host rock, where the calcareous cements of
the siltstones are easily dissolved by acids, leading
to the formation of the observed corrosion residues.
The large number of phylotypes representative of
Stenotrophomonas and Delftia spp. identied at sF88
(Table 2) is less easily explained, as members of these
genera carry out denitrication reactions, with the
conversion of ammonia to nitrous oxide (Dworkin,
2002). These organisms also play an important role
in the denitrication of complex organic compounds,
such as nitrobenzene. Such denitrication activity
Fig.6. X-ray powder diffractometry (XRD) of the samples at WF1
and sF88. The samples were segregated into an interior (bedrock)
and outer layer for a comparative analysis of mineralogical changes.
The analysis at WF1 reveals the predominantly calcite nature of
the bedrock, with an accumulation of clay particles in the surface
layer. At sF88 there is a dramatic change in the mineral structure,
with removal of the soluble dolomite and an enrichment of insoluble
silicates. Peaks correspond to: C = calcite; Q = quartz; D = dolomite; *
= unconsolidated endolite peak; and G = signature of glass support.
International Journal of Speleology, 36 (2), 93-104. Bologna (Italy). July 2007
Hazel A. Barton, Nicholas M. Taylor, Michael P. Kreate, Austin C. Springer, Stuart A. Oehrle and Janet L. Bertog.
101
is difcult to explain in the context of nitrogen
starvation, unless a key energy conservation activity
is nitrate reduction and/or the reduction of nitrogen-
containing aromatic compounds.
One interesting observation through the SEM-
EDX analyses was the selective enrichment of iron
oxides within the corrosion residues observed at
sF88, while our clone library does not demonstrate
the presence of any ‘classic’ iron-oxidizing species
(Table 2). One explanation may be that the iron-
oxidation that is observed on the surface of the rocks
could be the direct result of autoxidation, wherein
reduced iron within the host rock is exposed to the
oxygenated atmosphere of the cave through microbial
processes (Ehrlich, 2002). Nonetheless, it would be
surprising if the oligotrophic community at sF88 did
not harness Fe(II) as an electron donor before its loss.
The absence of well-known iron-oxidizing species may
reect the need for a more exhaustive phylogenetic
examination of this site (Ehrlich, 2002; Ley et al.,
2006). As with the WF1 community, nitrogen and
phosphorous must be growth limiting factors at sF88.
It is then hardly surprising that phylotypes related
to the nitrogen assimilating species Herbaspirillum
frisingense and Janthinobacterium agaricidamnosusm
were found. Interestingly Acinetobacter spp., which
can also mobilize inorganic phosphate, were identied
at sF88 and may provide an important clue for
nutrient acquisition in these starved ecosystems (Van
Groenestijn et al., 1988).
In attempting to understand the mechanisms
that support the often surprising levels of microbial
diversity in very starved cave environments, our
results suggest that community structure may be
greatly affected by the chemical nature of the rock
on which these organisms grow. In the case of the
WF1 community, which grows on limestone, the rock
has little potential for additional energy sources. As
a result, the community appears to rely more heavily
on heterotrophic growth from allochthonous energy
sources. The clay particles seen with XRD at this
site may be due to the production of organic acids by
microbial species, utilizing these reduced compounds
for growth, which leads to the accumulation of
these insoluble particles. At sF88, the geochemical
complexity of the rock may provide additional energy
sources, allowing species to use chemolithotrophic
mechanisms to conserve energy. The trace elements
available at sF88 could also prove essential to the
growth of microbial species, allowing the formation
of co-enzymes critical in intermediate metabolism.
Indeed, we have known for decades that many cell
types cannot grow without the addition of specic
mineral supplements (Conway de Macario et al.,
1982; Morgan, 1958; Roth et al., 1996). It is therefore
no surprise that the geochemistry of the bedrock can
impact both the microbial species capable of growth
as well as the types of energy conservation reactions
observed. The necessity for trace elements and
inorganic energy sources in growth is apparent at the
sF88 site, where microbial metabolic transformation
has led to extensive mineralogic alterations of the
Yates rock fabric and the formation of a corrosion
residue. Our results suggest that such variations in
geochemistry may have a profound affect on microbial
community structure in cave environments. As a
result, care should be taken when choosing sample
sites for microbial study within caves, as the geologic
setting may add unforeseen complexity to analyses or
complicate the interpretation of comparative studies.
Not only does this study hint at the high microbial
diversity in caves, in which niche biogeochemistry
may be an important driver of species diversity
(Begon et al., 1998), it also emphasizes the need for
a thorough understanding of the geologic conditions
when studying such environments.
ACKNOWLEDGEMENTS
The authors wish to thank Brad Lubbers for
excellent technical assistance, Karl Hagglund and
Brenda Racke for assistance with the SEM and
EDX analyses, Matthew Zacate for assistance in
establishing the ARB database and running the PAUP
software, Michael Queen for excellent assistance in
interpretation of the geology of Carlsbad Caverns,
and Harvey DuChene and an anonymous reviewer
for critical comments that signicantly improved the
manuscript. We would also like to thank the staff, in
particular Paul Burger, at the Cave Resource Ofce at
Carlsbad Caverns National Park for their invaluable
assistance with sample collection.
This work was supported in part by the Kentucky
EPSCoR Program, the Kentucky Academy of Science,
the Center for Integrative Natural Science and
Mathematics (CINSAM) at NKU, and the National Park
Service. Infrastructure support was provided, in part,
by the National Institutes of Heath KY INBRE program
(5P20RR016481-05). NMT and MPK were additionally
supported by NKU SURG awards.
REFERENCES
Achenbach L.A. & Coates J.D., 2000 - Disparity between
bacterial phylogeny and physiology. ASM News,
66: 714-715.
Altschul S.F., Madden T.L., Schaffer A.J., Zhang J., Zhang Z.,
Miller W. & Lipman D.J., 1997 - Gapped BLAST and PSI-
BLAST: a new generation of protein databae search programs.
Nuc. Acids Res., 25: 3389-3402.
Angert E.R., Northup D.E., Reysenbach A.-L., Peek A.S., Goebel
B.M. & Pace N.R., 1998 - Molecular phylogenetic analysis of a
bacterial community in Sulphur River, Parker Cave, Kentucky.
Am. Mineral., 83: 1583-1592.
Baneld J.F. & Nealson K.H., 1997 - Geomicrobiology: Interactions
between microbes and minerals. Rev. Mineral., Washington,
D.C. Mineralogical Society of American, 448 p.
Barton H.A. & Jurado V., 2007 - What’s Up Down There: Microbial
Diversity in Starved Cave Environments. Microbe, 2: 132-138.
Barton H.A. & Luiszer F., 2005 - Microbial Metabolic Structure
in a Suldic Cave Hot Spring: Potential Mechanisms of
Biospeleogenesis. J. Cave Karst Stud., 67: 28-38.
Barton H.A., Taylor M.R. & Pace N.R., 2004 - Molecular
Phylogenetic Analysis of a Bacterial Community in
an Oligotrophic Cave Environment. Geomicrobiol. J.,
21: 11-20.
The Impact of Host Rock Geochemistry on Bacterial Community Structure in Oligotrophic Cave Environments.
International Journal of Speleology, 36(2), 93-104. Bologna (Italy). July 2007
102
Sample Site
WF1 sF88
Element Average Wt%aSD Average Wt% a SD
C 12.32 ± 2.79 ND -
O 41.25 ± 0.72 49.11 ± 12.32
Mg 9.43 ± 0.31 1.18 ± 0.62
Ca 36.16 ± 4.44 ND -
Si 0.84 ± 0.64 33.55 ± 15.21
Al ND - 7.74 ± 3.98
K ND - 2.66 ± 1.23
Fe ND - 5.77 ± 4.25
Total 100% 100%
ND=none detected a=the average was taken from two separate analysis for 50 seconds at 2 kV (take of angle 38,3° count min-1
Table 1. EDX Bulk Elemental Analyses at WF1 and sF88
Phylogenetic Group Clone Clones
Group Closest identied relative % Sequence ID NCBI Accession
Number
sF88
Alphaproteobacteria NMTsF8 26/32 Brevundimonas nasdae 99% DQ066606
NMTsF27 5/32 Brevundimonas vesicularis 94% DQ066612
NMTsF32 1/32 Caulobacter subvibrioides 99% DQ066613
Betaproteobacteria NMTsF1 18/29 Massilia sp. 98% DQ066600
NMTsF3 6/29 Deltia tsuruhatasis 99% DQ066602
NMTsF7 2/29 Acidovorax sp. 98% DQ066605
NMTsF13 2/29 Herbaspirillum frisingense 100% DQ066607
NMTsF44 1/29 Janthinobacterium agaricidamnosusm 99% DQ066617
Gammaproteobacteria NMTsF2 17/31 Stenotrophomonas sp. 98% DQ066601
NMTsF6 1/31 Uncultured Acinetobacter sp. 99% DQ066604
NMTsF16 1/31 Uncultured bacteria clone FS117-02 89% DQ066608
NMTsF19 7/31 Cellvibrio ostraviensis 98% DQ066609
NMTsF26 4/31 Xanthomas retroexus 99% DQ066611
NMTsF36 1/13 Pseudomonas borealis 99% DQ066614
Cytophagales NMTsF40 1/1 Uncultured Bacteroidetes bacterium 98% DQ066616
Actinobacteria NMTsF4 2/7 Nocardioides sp. 98% DQ066603
NMTsF23 2/7 Rhodococcus sp. 100% DQ066610
NMTsF38 1/7 Mycobacterium gordonae 99% DQ066615
NMTsF47 2/7 Curtobacterium sp. 99% DQ066618
WF1 Library
Alphaproteobacteria NMT-WF15 1/1 Methylobacterium aquaticum 99% DQ228717
Betaproteobacteria NMT-WF23 1/1 Comamonas sp. 98% DQ228718
Gammaproteobacteria NMT-WF13 11/13 Uncultured bacterial mud-clone 90% DQ228716
NMT-WF31 2/13 Acinetobacter johnsonii 99% DQ228719
Actinobacteria NMT-WF1 34/65 Pseudonocardia sp. 97% DQ228711
NMT-WF4 14/65 Bacterium Chibacore 1500 90% DQ228712
NMT-WF7 4/65 Crossiella equi 97% DQ228713
NMT-WF9 1/65 Saccharothrix cryophillis 97% DQ228714
NMT-WF10 11/65 Actinomyces sp. 91% DQ228715
NMT-WF33 1/65 Actinobacterium sp. 93% DQ228720
Phylogenetic Group Clone Clones
Group Closest identied relative % Sequence ID NCBI Accession
Number
sF88
Alphaproteobacteria NMTsF8 26/32 Brevundimonas nasdae 99% DQ066606
NMTsF27 5/32 Brevundimonas vesicularis 94% DQ066612
NMTsF32 1/32 Caulobacter subvibrioides 99% DQ066613
Betaproteobacteria NMTsF1 18/29 Massilia sp. 98% DQ066600
NMTsF3 6/29 Deltia tsuruhatasis 99% DQ066602
NMTsF7 2/29 Acidovorax sp. 98% DQ066605
NMTsF13 2/29 Herbaspirillum frisingense 100% DQ066607
NMTsF44 1/29 Janthinobacterium agaricidamnosusm 99% DQ066617
Gammaproteobacteria NMTsF2 17/31 Stenotrophomonas sp. 98% DQ066601
NMTsF6 1/31 Uncultured Acinetobacter sp. 99% DQ066604
NMTsF16 1/31 Uncultured bacteria clone FS117-02 89% DQ066608
NMTsF19 7/31 Cellvibrio ostraviensis 98% DQ066609
NMTsF26 4/31 Xanthomas retroexus 99% DQ066611
NMTsF36 1/13 Pseudomonas borealis 99% DQ066614
Cytophagales NMTsF40 1/1 Uncultured Bacteroidetes bacterium 98% DQ066616
Actinobacteria NMTsF4 2/7 Nocardioides sp. 98% DQ066603
NMTsF23 2/7 Rhodococcus sp. 100% DQ066610
NMTsF38 1/7 Mycobacterium gordonae 99% DQ066615
NMTsF47 2/7 Curtobacterium sp. 99% DQ066618
WF1 Library
Alphaproteobacteria NMT-WF15 1/1 Methylobacterium aquaticum 99% DQ228717
Betaproteobacteria NMT-WF23 1/1 Comamonas sp. 98% DQ228718
Gammaproteobacteria NMT-WF13 11/13 Uncultured bacterial mud-clone 90% DQ228716
NMT-WF31 2/13 Acinetobacter johnsonii 99% DQ228719
Actinobacteria NMT-WF1 34/65 Pseudonocardia sp. 97% DQ228711
NMT-WF4 14/65 Bacterium Chibacore 1500 90% DQ228712
NMT-WF7 4/65 Crossiella equi 97% DQ228713
NMT-WF9 1/65 Saccharothrix cryophillis 97% DQ228714
NMT-WF10 11/65 Actinomyces sp. 91% DQ228715
NMT-WF33 1/65 Actinobacterium sp. 93% DQ228720
Table 2. Summary of the unique phylotype groups identied in the sF88 and WF1 clone libraries.
aSequences were compared against the NCBI GenBank database using a standard BLAST search (08/04; Altschul et al. 1997)
International Journal of Speleology, 36 (2), 93-104. Bologna (Italy). July 2007
Hazel A. Barton, Nicholas M. Taylor, Michael P. Kreate, Austin C. Springer, Stuart A. Oehrle and Janet L. Bertog.
103
Barton H.A., Taylor N.M., Lubbers B.R. & Pemberton A.C.,
2006 - DNA extraction from low-biomass carbonate rock:
an improved method with reduced contamination and
the low-biomass contaminant database. J. Microbiol.
Meth., 66: 21-31.
Begon M., Harper J.L. & Townsend C.R., 1998 - Ecology:
individuals, populations and communities. Cambridge,
Mass., Blackwell Scientic Publications, 1068 p.
Borer J.M. & Harris P.M., 1991 - Lithofacies and Cyclicity
of the Yates Formation, Permian Basin: Implications
for Reservoir Heterogeneity. Am. Assoc. Petrol. Geol.
Bull., 75: 726-779.
Boston P.J., Spilde M.N., Northup D.E., Melim L.A.,
Soroka D.S., Kleina L.G., Lavoie K.H., Hose L.D.,
Mallory L.M., Dahm C.N., Crossey L.J. & Schelble R.T.,
2001 - Cave Biosignature Suite: Microbes, Minerals
and Mars. Astrobiol., 1: 25-55.
Bottrell S.H., Smart P.L., Whitaker F. & Raiswell R.,
1991 - Geochemistry and isotope systematics of
sulphur in the mixing zone of Bahamian blue holes.
Appl. Geochem., 6: 97-103.
Brown A.A. & Loucks R.G., 1993 - Inuence of Sediment
Type and Depositional Processes on Stratal Patterns in
the Permian Basin-Margin Lamar Limestone, McKittrick
Canyon, Texas. Am. Assoc. Petrol. Geol. Bull.,
57: 435-474.
Canaveras J.C., Sanchez-Moral S., Soler V. & Saiz-
Jimenez C., 2001 - Microorganisms and microbially
induced fabrics in cave walls. Geomicrobiol. J.,
18: 223-240.
Chao A., 1987 - Estimating the population size for
capture-recapture data with unequal catchability.
Biometrics, 43: 783-791.
Chao A., Chazdon R.L., Colwell R.K. & Shen T.-J., 2005
- A new statistical approach for assembling similarity
of species composition with incidence and abundance
data. Ecol. Lett., 8: 148-159.
Chelius M.K. & Moore J.C., 2004 - Molecular
Phylogenetic Analysis of Archaea and Bacteria in
Wind Cave, South Dakota. Geomicrobiol. J., 21: 123-
134.
Colwell R.K., 2005, EstimateS: Statistical estimation of
species richness and shared species from samples.
Version 6.01b http://viceroy.eeb.uconn.edu/
estimates
Conway de Macario E., Macario A.J. & Wolin M.J., 1982
- Specic antisera and immunological procedures
for characterization of methanogenic bacteria. J.
Bacteriol., 149: 320-328.
Culver D.C., 1982 - Cave Life: Evolution and Ecology.
Cambridge, MA, Harvard University Press, 200 p.
DuChene H.R., 2000 - Bedrock Features of Lechuguilla
Cave, Guadalupe Mountains, New Mexico. J. Cave
Karst Stud., 62: 109-119.
Dworkin M., 2002 - The Prokaryotes: An evolving
electronic resource for the microbiological community.,
Springer-Verlag, New York.
Ehrlich H.L., 2002 - Geomicrobiology: New York, Marcel
Dekker, Inc., 768 p.
Engel A.S., Stern L.A. & Bennett P.C., 2004 - Microbial
contributions to cave formation: New insights into
sulfuric acid speleogenesis. Geol., 32: 369-372.
Forbes J.R., 2000 - Geochemistry of Carlsbad Cavern
Pool Waters, Guadalupe Mountains, New Mexico. J.
Cave Karst Stud., 62: 127-134.
Gotelli N.J. & Colwell R.K., 2001 - Quantifying
biodiversity: proceedures and pitfalls in the measurement
and comparison of species richness. Ecol. Lett.,
4: 379-391.
Grotenhuis J.T.C., Smit M., Plugge C.M., Yuansheng
X., Van Lammeren A.A.M., Stams A.J.M. & Zehnder
A.J.B., 1991 - Bacteriological composition and structure
of granular sludge adapted to different substrates. Appl.
Environ. Microbiol., 57: 1942-1949.
Groth I. & Saiz-Jimenez C., 1999 - Actinomycetes in
Hypogean Environments. Geomicrobiol. J., 16: 1-8.
Groth I., Schumann P., Laiz L., Moral-Sanchez
S., Canaveras J.C. & Saiz-Jimenez C., 2001 -
Geomicrobiological study of the Grotta dei Cervi, Porto
Badisco, Italy. Geomicrobiol. J., 18: 241-258.
Groth I., Vettermann R., Schuetze B., Schumann P.
& Saiz-Jimenez C., 1999 - Actinomycetes in Karstic
caves of Northern Spain (Altamira and Tito Bustillo). J.
Microbiol. Meth., 36: 115-122.
Hahn M.W. & Hoe M.G., 2001 - Grazing of protozoa
and its effect on populations of aquatic bacteria. FEMS
Microbiol. Ecol., 35: 113-121.
Hill C.A., 1990 - Sulfuric acid speleogenesis of Carlsbad
Cavern and its relationship to hydrocarbons, Delaware
Basin, New Mexico and Texas. Am. Assoc. Petrol. Geol.
Bull., 74: 1685-1694.
Hughes J.B., Hellmann J.J., Ricketts T.H. & Bohannan
B.J.M., 2001 - Counting the uncountable: Statistical
approaches to estimating microbial diversity. Appl.
Environ. Microbiol., 67: 4399-4406.
Itoh H. & Shiba T., 2004 - Polyphosphate synthetic
activity of polyphosphate:AMP phosphotransferase
in Actinetobacter johnsonii 210A. J. Bacteriol.,
186: 5178-5181.
Juttner F., 1984 - Dynamics of the volatile organic
substances associated with Cyanobacteria and Algae
in a eutrophic shallow lake. Appl. Environ. Microbiol.,
47: 814-820.
Klimchouk A.B., Ford D.C., Palmer A.N. & Dreybrodt W., 2000
- Speleogenesis: Evolution of Karstic Aquifers. Huntsville,
AL., National Speleological Society, 528 p.
Laiz L., Groth I., Gonzalez I. & Saiz-Jimenez C., 1999 -
Microbiological study of the dripping water in Altamira Cave
(Santillana del Mar, Spain). J. Microbiol. Meth., 36: 129-
138.
Ley R.E., Harris J.K., Wilcox J., Spear J.R., Miller S.R., Bebout
B.M., Maresca J.A., Bryant D.A., Sogin M.L. & Pace N.R.,
2006 - Unexpected diversity and complexity of the Guerrero
Negro hypersaline microbial mat. Appl. Environ. Microbiol.,
72: 3685-95.
Li Y., Kawamura Y., Fujiwara N., Naka T., Liu H., Huang X.,
Kobayashi K. & Ezaki T., 2004 - Sphingomonas yabuuchiae
sp. nov. and Brevundimonas nasdae sp. nov., isolated
from the Russian space laboratory Mir. Int. J. Syst. Evol.
Microbiol., 54: 819-825.
Maidak B.L., Cole J.R., Lilburn T.G., Parker C.T., Saxman P.R.,
Stredwick J.M., Garrity G.M., Li B., Olsen G.J., Pramanik
S., Schmidt T.M. & Tiedje J.M., 2000 - The RDP (Ribosomal
Database Project) continues. Nuc. Acids Res., 28: 173-174.
Phylogenetic Group Clone Clones
Group Closest identied relative % Sequence ID NCBI Accession
Number
sF88
Alphaproteobacteria NMTsF8 26/32 Brevundimonas nasdae 99% DQ066606
NMTsF27 5/32 Brevundimonas vesicularis 94% DQ066612
NMTsF32 1/32 Caulobacter subvibrioides 99% DQ066613
Betaproteobacteria NMTsF1 18/29 Massilia sp. 98% DQ066600
NMTsF3 6/29 Deltia tsuruhatasis 99% DQ066602
NMTsF7 2/29 Acidovorax sp. 98% DQ066605
NMTsF13 2/29 Herbaspirillum frisingense 100% DQ066607
NMTsF44 1/29 Janthinobacterium agaricidamnosusm 99% DQ066617
Gammaproteobacteria NMTsF2 17/31 Stenotrophomonas sp. 98% DQ066601
NMTsF6 1/31 Uncultured Acinetobacter sp. 99% DQ066604
NMTsF16 1/31 Uncultured bacteria clone FS117-02 89% DQ066608
NMTsF19 7/31 Cellvibrio ostraviensis 98% DQ066609
NMTsF26 4/31 Xanthomas retroexus 99% DQ066611
NMTsF36 1/13 Pseudomonas borealis 99% DQ066614
Cytophagales NMTsF40 1/1 Uncultured Bacteroidetes bacterium 98% DQ066616
Actinobacteria NMTsF4 2/7 Nocardioides sp. 98% DQ066603
NMTsF23 2/7 Rhodococcus sp. 100% DQ066610
NMTsF38 1/7 Mycobacterium gordonae 99% DQ066615
NMTsF47 2/7 Curtobacterium sp. 99% DQ066618
WF1 Library
Alphaproteobacteria NMT-WF15 1/1 Methylobacterium aquaticum 99% DQ228717
Betaproteobacteria NMT-WF23 1/1 Comamonas sp. 98% DQ228718
Gammaproteobacteria NMT-WF13 11/13 Uncultured bacterial mud-clone 90% DQ228716
NMT-WF31 2/13 Acinetobacter johnsonii 99% DQ228719
Actinobacteria NMT-WF1 34/65 Pseudonocardia sp. 97% DQ228711
NMT-WF4 14/65 Bacterium Chibacore 1500 90% DQ228712
NMT-WF7 4/65 Crossiella equi 97% DQ228713
NMT-WF9 1/65 Saccharothrix cryophillis 97% DQ228714
NMT-WF10 11/65 Actinomyces sp. 91% DQ228715
NMT-WF33 1/65 Actinobacterium sp. 93% DQ228720
Phylogenetic Group Clone Clones
Group Closest identied relative % Sequence ID NCBI Accession
Number
sF88
Alphaproteobacteria NMTsF8 26/32 Brevundimonas nasdae 99% DQ066606
NMTsF27 5/32 Brevundimonas vesicularis 94% DQ066612
NMTsF32 1/32 Caulobacter subvibrioides 99% DQ066613
Betaproteobacteria NMTsF1 18/29 Massilia sp. 98% DQ066600
NMTsF3 6/29 Deltia tsuruhatasis 99% DQ066602
NMTsF7 2/29 Acidovorax sp. 98% DQ066605
NMTsF13 2/29 Herbaspirillum frisingense 100% DQ066607
NMTsF44 1/29 Janthinobacterium agaricidamnosusm 99% DQ066617
Gammaproteobacteria NMTsF2 17/31 Stenotrophomonas sp. 98% DQ066601
NMTsF6 1/31 Uncultured Acinetobacter sp. 99% DQ066604
NMTsF16 1/31 Uncultured bacteria clone FS117-02 89% DQ066608
NMTsF19 7/31 Cellvibrio ostraviensis 98% DQ066609
NMTsF26 4/31 Xanthomas retroexus 99% DQ066611
NMTsF36 1/13 Pseudomonas borealis 99% DQ066614
Cytophagales NMTsF40 1/1 Uncultured Bacteroidetes bacterium 98% DQ066616
Actinobacteria NMTsF4 2/7 Nocardioides sp. 98% DQ066603
NMTsF23 2/7 Rhodococcus sp. 100% DQ066610
NMTsF38 1/7 Mycobacterium gordonae 99% DQ066615
NMTsF47 2/7 Curtobacterium sp. 99% DQ066618
WF1 Library
Alphaproteobacteria NMT-WF15 1/1 Methylobacterium aquaticum 99% DQ228717
Betaproteobacteria NMT-WF23 1/1 Comamonas sp. 98% DQ228718
Gammaproteobacteria NMT-WF13 11/13 Uncultured bacterial mud-clone 90% DQ228716
NMT-WF31 2/13 Acinetobacter johnsonii 99% DQ228719
Actinobacteria NMT-WF1 34/65 Pseudonocardia sp. 97% DQ228711
NMT-WF4 14/65 Bacterium Chibacore 1500 90% DQ228712
NMT-WF7 4/65 Crossiella equi 97% DQ228713
NMT-WF9 1/65 Saccharothrix cryophillis 97% DQ228714
NMT-WF10 11/65 Actinomyces sp. 91% DQ228715
NMT-WF33 1/65 Actinobacterium sp. 93% DQ228720
The Impact of Host Rock Geochemistry on Bacterial Community Structure in Oligotrophic Cave Environments.
International Journal of Speleology, 36(2), 93-104. Bologna (Italy). July 2007
104
Mohn W.W. & Tiedje J.M., 1992 - Microbial reductive
dehalogenation. Microbiol. Rev., 56: 482-507.
Morgan J.F., 1958 - Tissue Culture Nutrition. Bacteriol.
Rev., 22: 20-45.
Mutti M. & Simo J.A., 1993 - Stratigraphic Patterns and
Cycle-Related Diagenesis of Upper Yates Formation,
Permian, Guadalupe Mountains. Am. Assoc. Petrol.
Geol. Bull., 57: 515-534.
Nei M., Kumar S. & Takahashi K., 1998 - The optimization
principle in phylogenetic analysis tends to give incorret
topologies when the number of nucleotide or amino acids
used is small. Proc. Ntl. Acad. Sci. USA., 95: 12390-
12397.
Northup D.E., Barnes S.M., Yu L.E., Spilde M.N., Schelble
R.T., Dano K.E., Crossey L.J., Connolly C.A., Boston
P.J., Natvig D.O. & Dahm C.N., 2003 - Diverse microbial
communitiens inhabiting ferromanganese deposits in
Lechuguilla and Spider Caves. Environ. Microbiol.,
5: 1071-1086.
Osman S., Stuecker T., Newcombe D. & Venkateswaran
K., 2005 - Microbial burden and community proles of
the genesis curation laboratory. Abstract 105th General
Meeting, Am. Soc. Microbiol., Atlanta, GA.
Pace N.R., 1997 - A molecular view of microbial diversity
and the biosphere. Science, 276: 734-740.
Palmer A.N., 2000 - Hydrochemical Interpretation of Cave
Patterns in the Guadalupe Mountains, New Mexico. J.
Cave Karst Stud., 62: 91-108.
Roth J.R., Lawrence J.G. & Bobik T.A., 1996 - Cobalamin
(coenzyme B12): synthesis and biological signicance.
Ann. Rev. Microbiol., 50: 137-181.
Saiz-Jimenez C. & Hermosin B., 1999 - Thermally asisted
hydrolysis and methylation of dissolved organic matter
in dripping waters from the Altamira Cave. J. Anal.
Appl. Pyrol., 49: 337-347.
Sarbu S.M., Kane T.C. & Kinkle B.K., 1996 - A
chemoautotrophically based cave ecosystem. Science,
272: 1953-1955.
Schink B., 2002 - Synergistic interactions in the microbial
world. Anton. Leeuw., 81: 257-261.
Sylvia D.M., Fuhrmann J.J., Hartel P.G. & Zuberer D.A.,
1999 - Principles and Applications of Soil Microbiology.
Upper Saddle River, NJ, Prentice Hall, 550 p.
Van Groenestijn J.W., Vlekke G.J.F.M., Anink D.M.E.,
Deinema M.H. & Zehnder A.J.B., 1988 - Role of
cations in accumulation and release of phosphate by
Acinetobacter strain 210A. Appl. Environ. Microbiol.,
54: 2894-2901.
International Journal of Speleology, 36 (2), 93-104. Bologna (Italy). July 2007
Hazel A. Barton, Nicholas M. Taylor, Michael P. Kreate, Austin C. Springer, Stuart A. Oehrle and Janet L. Bertog.
... As cavernas são pouco estudadas do ponto de vista microbiológico e as características deste ambiente propiciam um habitat microbiano único, com pressões evolutivas diferentes do ambiente epígeo (Barton et al., 2007). Com isso, os organismos presentes são mais especializados ao baixo aporte energético das cavernas, o que aumenta o interesse na identificação de microrganismos e estudos de caracterização biotecnológica. ...
... Caves are scarcely studied from the microbiological point of view. The characteristics of this environment provide a unique microbial habitat, with different evolutionary pressures from the epigeal environment (Barton et al., 2007). The organisms present are more specialized to the low energy input of caves, which increases the interest in the identification of microorganisms and biotechnological characterization studies. ...
... Fatores como baixo custo de produção, alta produtividade, estabilidade a temperaturas extremas, especificidade, pH ou outras condições fisiológicas fazem com que o estudo de bactérias em ambientes peculiares, como o cavernícola, seja uma alternativa em destaque na busca por novos medicamentos (Prakash et al., 2013). As pressões seletivas do ambiente podem favorecer grupos de bactérias com capacidade de produção de antimicrobianos (Barton et al., 2007), por exemplo, as actinobactérias, as quais têm sido amplamente encontradas nos ambientes cavernícolas (Jurado et al., 2005). ...
Article
Full-text available
A Gruta Martimiano II é a maior caverna quartzítica do Brasil e está localizada no Parque Estadual do Ibitipoca – Minas Gerais. As condições ambientais únicas e o baixo aporte energético das cavidades atuam como pressões seletivas nos organismos que as habitam. A microbiota tem grande importância ecológica e biotecnológica, uma vez que microrganismos cultiváveis são amplamente utilizados como recursos na agricultura, indústria farmacêutica e em bioprocessos. Entretanto, são poucos os estudos sobre a microbiota cavernícola e seu potencial biotecnológico. Os objetivos deste trabalho foram bioprospectar e analisar o potencial biotecnológico de bactérias associadas ao piso, teto e parede da Gruta Martimiano II. Um total de 12 amostras oriundas de quatro áreas da caverna foram obtidas. A caracterização das fases minerais foi feita por difração de raios X. As amostras foram então lavadas e diluídas em série com água de torneira autoclavada. O lavado foi plaqueado em meio de cultura rico e após 3 dias as colônias bacterianas foram isoladas, preservadas, os testes bioquímicos foram realizados e o isolado de maior potencial foi identificado. As fases minerais caracterizadas foram as mesmas entre as amostras, compostas por quartzo, muscovita, caulinita e gibbsita. Foram obtidos 72 isolados bacterianos que constituem hoje primeiro banco de isolados bacterianos desta caverna. Três isolados tiveram a capacidade em fixar nitrogênio e um em solubilizar fosfato, componentes fundamentais ao desenvolvimento de plantas. Cinco isolados apresentaram possível antagonismo a Klebsiella pneumoniae e o isolado 14 foi capaz de inibir o crescimento de K. pneumoniae e Staphylococcus aureus. Em ensaio de inibição indireta o isolado 14 impediu o crescimento dos patógenos, foi caracterizado como móvel, Gram-negativo e identificado como Peribacillus. O ineditismo desta pesquisa corrobora o potencial de geração de novos conhecimentos acompanhado do desenvolvimento de novos produtos, agregando valor a serviços ecossistêmicos oriundos de um patrimônio genético desconhecido.
... Basalts are N-poor, generally less than 0.01% 172 . The addition of N to seafloor basalts in incubation experiments stimulated the growth of Fe-oxidizing taxa, which suggests that these groups are generally N-limited 173,174 . Additionally, multiple studies have shown that basalt can enhance the growth of organisms in Felimited media by providing structural Fe(II) for Fe-oxidation pathways 13,161,167 . ...
Article
Full-text available
Volcanic eruptions generate initially sterile materials where biological processes are absent, allowing for the fresh colonization by new organisms. This review summarizes the characteristics of volcanic habitats that are available for pioneer microbial colonization, including hot springs, fumaroles, lava tubes, and recently cooled rock surfaces and interiors. Eruptions provide unique insight into microbial community development in extreme environments. The trajectories that these ecosystems follow are largely dictated by the initial environmental conditions and identities of the colonizers, rather than the age of the system. The review also discusses how studies of microbial communities in young lava flow fields can provide insights into the possibility of life on Mars, which was volcanically and hydrologically active in the past. Understanding biosignature preservation as well as the metabolisms and survival mechanisms of microorganisms in volcanic systems has implications for how an ecosystem might have developed on early Earth and possibly Mars.
... For example, soil microbial communities (including bacteria and fungi) on limestone grouped in a significant cluster (Figure 7), even though the sites were at 1200 and 2100 m. This result can be explained by the differing soil conditions between limestone sites and clasolite sites, including soil pH, SOC concentration, and soil clay content, which corresponds with the fact that differences in mineral and element composition or soil texture are important in shaping microbial community composition (Barton et al., 2007;Tytgat et al., 2016). ...
Article
Aim Climate is widely understood to determine elevational patterns of soil microbial communities, whereas the effects of parental material are uncertain. Changes in the composition of parental materials along elevational transects could also affect soil microbial communities by influencing soil pH and nutrient availability. Here, we aim to illustrate the combined effects of climate and parental material on the biomass and composition of soil microbial communities along an elevational transect. Location A subtropical forest on a karst mountain (Mt. Jinfo), China. Taxon Bacteria and Fungi. Methods We use phospholipid fatty acid analysis (PLFA) and DNA amplicon high‐throughput sequencing to determine biomass and diversity patterns of soil microbial communities along a subtropical elevational gradient with contrasting parental materials (limestone and clasolite). Results We observed that the microbial communities were more diverse (α‐diversity) and productive (biomass) on limestone than on clasolite. Additionally, we found that parental material played a role in shaping the composition (β‐diversity) of soil microbial communities along the elevational gradient. The impact of climate on soil microbial communities was found to be significant, albeit relatively weak. Structural equation models provided evidence for both direct and indirect effects of climate and parental material on microbial biomass and α‐diversity along the elevational gradient. Notably, the changes in soil pH, influenced by both parental material and climate, were identified as a key factor driving these effects. Main Conclusions Our results underline the importance of both climate and parental material variations in space‐for‐time studies investigating soil microbial communities along elevational gradients.
... energy source) (Castle et al. 2017 ). The main factors that determine nutrient availability of newly exposed oligotrophic polar soils remain largely unexplored but are pr esumabl y r elated to soil bedr oc k type (Barton et al. 2007, Lazzaro et al. 2009, Tytgat et al. 2016, Wojcik et al. 2019 ) and/or soil organic matter sources (Fierer et al. 2010 ). ...
Article
Full-text available
Substrate geochemistry is an important factor influencing early microbial development after glacial retreat on nutrient-poor geological substrates in the High Arctic. It is often difficult to separate substrate influence from climate because study locations are distant. Our study in the retreating Nordenskiöldbreen (Svalbard) is one of the few to investigate biogeochemical and microbial succession in two adjacent forefields, which share the same climatic conditions but differ in their underlying geology. The northern silicate forefield evolved in a classical chronosequence, where most geochemical and microbial parameters increased gradually with time. In contrast, the southern carbonate forefield exhibited high levels of nutrients and microbial biomass at the youngest sites, followed by a significant decline and then a gradual increase, which caused a rearrangement in the species and functional composition of the bacterial and fungal communities. This shuffling in the early stages of succession suggests that high nutrient availability in the bedrock could have accelerated early soil succession after deglaciation and thereby promoted more rapid stabilization of the soil and production of higher quality organic matter. Most chemical parameters and bacterial taxa converged with time, while fungi showed no clear pattern.
... The Trenton layer community was explained by Nocardioides and Limnobacter, and Lysinimicrobium, Tabrizicola, Thalassospira, and Oxalicibacterium were unique taxa. Nocardioides is found in terrestrial and aquatic environments, as well as low-nutrient habitats such as oligotrophic cave rocks, marine sediments, or Siberian permafrost [67][68][69]. Lysinimicrobium and Tabrizicola are both tolerant to high salt concentrations [70,71], possible remnants of the marine origin of the rocks. Limnobacter and Thalassospira both exhibit metabolisms useful in low-nutrient oligotrophic conditions (thiosulfate oxidation [72], phosphate chemotaxis [73]). ...
Article
Full-text available
The deep terrestrial subsurface, hundreds of meters to kilometers below the surface, is characterized by oligotrophic conditions, dark and often anoxic settings, with fluctuating pH, salinity, and water availability. Despite this, microbial populations are detected and active, contributing to biogeochemical cycles over geological time. Because it is extremely difficult to access the deep biosphere, little is known about the identity and metabolisms of these communities, although they likely possess unknown pathways and might interfere with deep waste deposits. Therefore, we analyzed rock and groundwater microbial communities from deep, isolated brine aquifers in two regions dating back to the Ordovician and Devonian, using amplicon and whole genome sequencing. We observed significant differences in diversity and community structure between both regions, suggesting an impact of site age and composition. The deep hypersaline groundwater did not contain typical halophilic bacteria, and genomes suggested pathways involved in protein and hydrocarbon degradation, and carbon fixation. We identified mainly one strategy to cope with osmotic stress: compatible solute uptake and biosynthesis. Finally, we detected many bacteriophage families, potentially indicating that bacteria are infected. However, we also found auxiliary metabolic genes in the viral genomes, probably conferring an advantage to the infected hosts.
... However, caves host high diverse microbes (Pedersen, 2000). In the meanwhile, the cave entrances mostly directly connected to the outside and is more easily accessible than other subsurface environment (Barton et al., 2007), thus karst caves are an ideal sites to study microbial diversity and function of the subsurface biosphere (Barton et al., 2004). ...
Article
There is limited knowledge about microbial communities and their ecological functions in karst caves with high CO2 concentrations. Here, we studied the microbial community compositions and functions in Shuiming Cave ("SMC", CO2 concentration 3303 ppm) and Xueyu Cave ("XYC", CO2 concentration 8753 ppm) using Illumina MiSeq high-throughput sequencing in combination with BIOLOG test. The results showed that Proteobacteria, Actinobacteria, and Bacteroidetes were dominant phyla in these two caves, and Thaumarchaeota was the most abundant in the rock wall samples of SMC. The microbial diversity in the water samples decreased with increasing HCO3- concentration, and it was higher in XYC than that in SMC. The microbial community structures in the sediment and rock wall samples were quite different between the two caves. High concentrations of CO2 can reduce the microbial diversity on the rock walls in karst caves, probably through changing microbial preference for different types of carbon sources and decreasing the microbial utilization rate of carbon sources. These results expanded our understanding of microbial community and its response to environments in karst caves with high CO2 .
... Most microorganisms in subsurface environments grow as biofilms or individual or groups of cells attached to rock surfaces. Rocks and minerals are not homogeneous and can influence microbial diversity at the microniche level [59,60] Components of the stratum may be toxic or beneficial to microbial growth by providing mineral nutrients, pH buffering, or other advantageous conditions. Given the long evolutionary history of microbes in geologic time, Jones and Bennett [60] hypothesize that each mineral surface is specifically altered by the best-adapted and most comprehensive microbial community that can use the mineral surface to the greatest advantage. ...
Article
Full-text available
The genus Crossiella contains two species, C. equi, causing nocardioform placentitis in horses, and C. cryophila, an environmental bacterium. Apart from C. equi, which is not discussed here, environmental Crossiella is rarely reported in the literature; thus, it has not been included among “rare actinobacteria”, whose isolation frequency is very low. After C. cryophila, only five reports cover the isolation of Crossiella strains. However, the frequency of published papers on environmental Crossiella has increased significantly in recent years due to the extensive use of next-generation sequencing (NGS) and a huge cascade of data that has improved our understanding of how bacteria occur in the environment. In the last five years, Crossiella has been found in different environments (caves, soils, plant rhizospheres, building stones, etc.). The high abundance of Crossiella in cave moonmilk indicates that this genus may have an active role in moonmilk formation, as evidenced by the precipitation of calcite, witherite, and struvite in different culture media. This review provides an overview of environmental Crossiella, particularly in caves, and discusses its role in biomineralization processes and bioactive compound production.
Article
Full-text available
Microbial communities inhabiting caves in quartz-rich rocks are still underexplored, despite their possible role in the silica cycle. The world's longest orthoquartzite cave, Imawarì Yeuta, represents a perfect arena for the investigation of the interactions between microorganisms and silica in non-thermal environments due to the presence of extraordinary amounts of amorphous silica speleothems of different kinds. In this work, the microbial diversity of Imawarì Yeuta was dissected by analyzing nineteen samples collected from different locations representative of different silica amorphization phases and types of samples. Specifically, we investigated the major ecological patterns in cave biodiversity, specific taxa enrichment, and the main ecological clusters through co-occurrence network analysis. Water content greatly contributed to the microbial communities' composition and structures in the cave leading to the sample clustering into three groups DRY, WET, and WATER. Each of these groups was enriched in members of Actinobacteriota, Acidobacteriota, and Gammaproteobacteria, respectively. Alpha diversity analysis showed the highest value of diversity and richness for the WET samples, while the DRY group had the lowest. This was accompanied by the presence of correlation patterns including either orders belonging to various phyla from WET samples or orders belonging to the Actinobacteriota and Firmicutes phyla from DRY group samples. The phylogenetic analysis of the dominant species in WET and DRY samples showed that Acidobacteriota and Actinobacteriota strains were affiliated with uncultured bacteria retrieved from various oligotrophic and silica/quartz-rich environments, not only associated with subterranean sites. Our results suggest that the water content greatly contributes to shaping TYPE Original Research
Article
Full-text available
Sulphur River in Parker Cave, Kentucky receives sulfurous water (11-21 mg sulfide/L) from the Phantom Waterfall and contains a microbial mat composed of white filaments. We extend a previous morphological survey with a molecular phylogenetic analysis of the bacteria of the microbial mat. This approach employs DNA sequence comparisons of small subunit ribosomal RNA (SSU rRNA) genes obtained from the mat with those from an extensive database of rRNA sequences. Many of SSU rRNA gene clones obtained from the mat are most similar to rRNA sequences from sulfur-oxidizing bacteria (Thiothrix spp., Thiomicrospira denitrificans, and "Candidatus Thiobacillus baregensis"). The Sulphur River SSU rRNA gene clones also show specific affiliations with clones from environmental surveys of bacteria from deep-sea hydrothermal vent communities and subsurface microcosms. Affiliations with sequences from bacteria that are known to have the ability to obtain energy for CO2 fixation from the oxidation of inorganic compounds (chemoautotrophs), in combination with the environmental conditions surrounding the microbial mat, indicate that chemoautotrophic metabolism of bacteria in this mat may contribute to the biomass of Sulphur River. Cave communities, such as the one identified in Sulphur River, provide sites to study such relatively autonomous chemoautotrophic communities that are much more accessible than similar communities associated with deep-sea hydrothermal vents. Subsurface microbiology and the contribution of microbial activity on cave development are also discussed.
Article
Full-text available
Most caves in the Guadalupe Mountains have ramifying patterns consisting of large rooms with narrow rifts extending downward, and with successive outlet passages arranged in crude levels. They were formed by sulfuric acid from the oxidation of hydrogen sulfide, a process that is now dormant. Episodic escape of H2S-rich water from the adjacent Delaware Basin, and perhaps also from strata beneath the Guadalupes, followed different routes at different times. For this reason, major rooms and passages correlate poorly between caves, and within large individual caves. The largest cave volumes formed where H2S emerged at the contemporary water table, where oxidation was most rapid. Steeply ascending passages formed where oxygenated meteoric water converged with deep-seated H2S-rich water at depths as much as 200 m below the water table. Spongework and network mazes were formed by highly aggressive water in mixing zones, and they commonly rim, underlie, or connect rooms. Transport of H2S in aqueous solution was the main mode of H2S influx. Neither upwelling of gas bubbles nor molecular diffusion appears to have played a major role in cave development, although some H2S could have been carried by less-soluble methane bubbles. Most cave origin was phreatic, although subaerial dissolution and gypsum-replacement of carbonate rock in acidic water films and drips account for considerable cave enlargement above the water table. Estimates of enlargement rates are complicated by gypsum replacement of carbonate rock because the gypsum continues to be dissolved by fresh vadose water long after the major carbonate dissolution has ceased. Volume-for-volume replacement of calcite by gypsum can take place at the moderate pH values typical of phreatic water in carbonates, preserving the original bedrock textures. At pHs less than about 6.4, this replacement usually takes place on a molar basis, with an approximately two-fold volume increase, forming blistered crusts.
Article
After several decades of microbiological research has focused on pure cultures, synergistic effects between different types of microorganisms find increasing interest. Interspecies interactions between prokaryotic cells have been studied into depth mainly with respect to syntrophic cooperations involved in methanogenic degradation of electron-rich substrates such as fatty acids, alcohols, and aromatics. Partners involved in these processes have to run their metabolism at minimal energy increments, with only fractions of an ATP unit synthesized per substrate molecule metabolized, and their cooperation is intensified by close proximity of the partner cells. New examples of such syntrophic activities are anaerobic methane oxidation by presumably methanogenic and sulfate-reducing prokaryotes, and microbially mediated pyrite formation. Syntrophic relationships have also been discovered to be involved in the anaerobic metabolization of amino acids and sugars where energetical restrictions do not necessarily force the partner organisms into strict interdependencies. The most highly developed cooperative systems among prokaryotic cells appear to be the structurally organized phototrophic consortia of the Chlorochromatium and Pelochromatium type in which phototrophic and chemotrophic bacteria not only exchange metabolites but also interact at the level of growth coordination and tactic behaviour.
Article
The sediment type delivered to the basin margin varied systematically during the Late Guadalupian due to sea-level fluctuations. Relative sea-level stand was interpreted from correlation to the relative sea-level record of contemporaneous shelf strata and from analogy to the Bahamian Quaternary carbonate sediment history. Siliciclastic silts were deposited in the basin during lowstands as onlapping strata. -from Authors
Article
Glenwood Hot Springs, Colorado, is a sulfidic hot-spring that issues from numerous sites. These waters are partially responsible for speleogenesis of the nearby Fairy Cave system, through hypogenic sulfuric-acid dissolution. To examine whether there may have been microbial involvement in the dissolution of this cave system we examined the present-day microbial flora of a cave created by the hot spring. Using molecular phylogenetic analysis of the 16S small subunit ribosomal RNA gene and scanning electron microscopy, we examined the microbial community structure within the spring. The microbial community displayed a high level of microbial diversity, with 25 unique phylotypes representing nine divisions of the Bacteria and a division of the Archaea previously not identified under the conditions of temperature and pH found in the spring. By determining a putative metabolic network for the microbial species found in the spring, it appears that the community is carrying out both sulfate reduction and sulfide oxidation. Significantly, the sulfate reduction in the spring appears to be generating numerous organic acids as well as reactive sulfur species, such as sulfite. Even in the absence of oxygen, this sulfite can interact with water directly to produce sulfuric acid. Consequently, such metabolic activity may represent a mechanism by which biospeleogenesis can lead to passage enlargement through sulfuric acid production without the influx of oxygen or oxygen-rich waters. Such activity may lead to higher levels of sulfuric acid production than could be accounted for by inorganic hydrogen sulfide oxidation. Therefore, rather than generating localized pockets of speleogenesis within cave systems, such biogenic sulfuric acid production may have a regional impact on water chemistry and subsequent speleogenesis of large cave systems.
Article
Water samples collected from 13 pools in Carlsbad Cavern were analyzed to determine the concentrations of major ions. Air temperature, relative humidity, and carbon dioxide concentration of the cave atmosphere were also measured. Large differences in water quality exist among different cave pools, with some pools containing very fresh water, while others are brackish, with total dissolved solids concentrations up to 5000 mg/L. Brackish water pools appear to be associated with those portions of the cave where evaporation rates are high and/or soluble minerals are present. Geochemical speciation modeling showed that some pools are close to saturation with respect to the common cave minerals aragonite, calcite, gypsum, and hydromagnesite. A tracer test was performed using a non-toxic bromide salt to estimate the leakage rates of selected pools. Pool volumes calculated based on dilution of the bromide tracer were up to 550 m3. The tracer test results were used to calculate mean residence times for the water in each pool. Calculated mean residence times based on bromide tracer loss rates ranged from less than a year for Rookery Pool and Devil's Spring to 16 years for Lake of the Clouds. Calculated pool leakage rates ranged from 2 L/day to over 100 L/day. The pools with the highest leakage rates appear to be Rookery Pool, Green Lake, and Lake of the Clouds. The long residence times indicated by the tracer tests suggest that the pools evaporate more water than they leak. However, evaporation should result in an accumulation of dissolved chloride and other solutes in the pools, which for most pools does not appear to be the case. Taken together, these observations suggest that the pools are recharged primarily by infrequent precipitation events, separated by long periods of slow evaporation and minimal leakage.
Article
Capitan backreef strata of the upper Yates and lowermost Tansill formations are characterized by three cycles composed of a lower siliciclastic unit and an upper carbonate unit. The cycles are bounded by sharp and erosive surfaces on the shelf that become concordant and disappear within concordant outer-shelf margin strata. Syndepositional diagenetic events record several phases of marine cementation, neomorphism and dissolution, replacive dolomitization, and primary dolomite precipitation. Composite sea-level variations controlled the deposition of cyclic strata and distribution of syndepositional diagenetic features. Because syndepositional diagenesis records the effects of relative eustatic variations, its distribution can be predicted. -from Authors