Subido por Héctor Galván

Azsalos et al 2016 Extremófilas

Anuncio
Extremophiles (2016) 20:603–620
DOI 10.1007/s00792-016-0849-3
ORIGINAL PAPER
Diversity of extremophilic bacteria in the sediment
of high‑altitude lakes located in the mountain
desert of Ojos del Salado volcano, Dry‑Andes
Júlia Margit Aszalós1 · Gergely Krett1 · Dóra Anda1 · Károly Márialigeti1 ·
Balázs Nagy2 · Andrea K. Borsodi1
Received: 15 March 2016 / Accepted: 31 May 2016 / Published online: 17 June 2016
© Springer Japan 2016
Abstract Ojos del Salado, the highest volcano on Earth is
surrounded by a special mountain desert with extreme aridity, great daily temperature range, intense solar radiation,
and permafrost from 5000 meters above sea level. Several
saline lakes and permafrost derived high-altitude lakes can
be found in this area, often surrounded by fumaroles and
hot springs. The aim of this study was to gain information about the bacterial communities inhabiting the sediment of high-altitude lakes of the Ojos del Salado region
located between 3770 and 6500 m. Altogether 11 sediment
samples from 4 different altitudes were examined with 16S
rRNA gene based denaturing gradient gel electrophoresis
and clone libraries. Members of 17 phyla or candidate divisions were detected with the dominance of Proteobacteria,
Acidobacteria, Actinobacteria and Bacteroidetes. The bacterial community composition was determined mainly by
the altitude of the sampling sites; nevertheless, the extreme
aridity and the active volcanism had a strong influence on
it. Most of the sequences showed the highest relation to
bacterial species or uncultured clones from similar extreme
environments.
Keywords High-altitude lakes · Dry-Andes · Bacterial
diversity · 16S rRNA gene · DGGE · Clone library
Communicated by A. Oren.
* Andrea K. Borsodi
[email protected]
1
Department of Microbiology, Eötvös Loránd University,
Pázmány Péter sétány 1/C, 1117 Budapest, Hungary
2
Department of Physical Geography, Eötvös Loránd
University, Pázmány P. sétány 1/C, 1117 Budapest, Hungary
Introduction
Ojos del Salado (6893 m, S 27°06′34.6″, W 68°32′32.1″),
the highest volcano on Earth is a dormant stratovolcano
hosted by the deserted plateau of Puna de Atacama (with
an average altitude of 4500 meters above sea level; m.a.s.l)
in the Dry-Andes. As an effect of the South American Dry
Diagonal, the plateau is characterized by extreme aridity (lack of glaciers and increased elevation of potential
snowline altitude), forming a remote mountain desert up to
6000 m (Vuille and Ammann 1997; Ammann et al. 2001;
Houston and Hartley 2003; Azócar and Brenning 2010;
Nagy et al. 2014a).
Mountain deserts are harsh environments characterized by extreme aridity, intense solar UV radiation and
extreme shifts in daily temperature (even 50 °C on the
surface). In the region of Ojos del Salado precipitation
occurs sporadically, and the snow rapidly sublimates
instead of melting (Nagy et al. 2014a). In this hyperarid environment, altitudinal limit of vascular plants and
continuous vegetation is 4600 m.a.s.l., much lower than
in the case of the more humid Himalaya (6350 m.a.s.l.)
(Halloy 1991).
In mountain deserts water is present mainly in the form
of ice within permafrost. Permafrost by definition of the
Permafrost Subcommittee (1988) is “ground (i.e., soil and
rock) that remains at or below 0 °C for at least 2 years”.
In the region of Ojos del Salado from 5000 m.a.s.l discontinuous, while from 5600 m.a.s.l continuous permafrost can
be found. The active layer of permafrost thaws periodically
as a result of temperature changes. Although, the melting
affects only the upper 5–60 cm of the permafrost, the active
layer slowly increases as an effect of global climate change
(Nagy et al. 2014b). Meltwater from thawing permafrost
frequently accumulates in endorheic basins forming small
13
604
and shallow high-altitude lakes. These permafrost originated lakes are among the highest altitude lakes on Earth.
Nevertheless, in a longer period of time they might disappear due to the degradation of permafrost and desiccation
of the regolith (Nagy et al. 2014a, b).
Beside lakes derived from the degrading permafrost,
several salt-flats and high-altitude saline lakes (e.g. Laguna
Santa Rosa, Laguna Verde) can also be found in the Puna
de Atacama. The volumes of these high-altitude DryAndean lakes are decreasing since the last humid period
because evaporation outweighs the amount of precipitation
(Demergasso et al. 2010).
The high-altitude lakes of Ojos del Salado are extreme
environments resulting from the properties of mountain
deserts, including intense solar radiation, radical changes in
daily temperature and low water activity. Volcanic activities
also have an influence on the physical and chemical characteristics of the environment; several hot springs are present in this area which can alter the pH, and—along with
the solar radiation—affect the water temperature in a wide
range. Fumaroles in high-altitude sites have been reported
to host diverse communities with mosses, lichens, fungi
and microbial mats in otherwise blank places (Halloy 1991;
Costello et al. 2009).
High-altitude lakes are sensitive ecosystems which
respond rapidly to environmental changes; therefore, they
are natural laboratories and indicators for climate changes
(Adrian et al. 2009). They also give us insights into early
life on Earth when the lack of ozone layer caused intense
UV radiation and extremities in temperature (Cabrol et al.
2007). These lakes might be inhabited by several hitherto
unknown microorganisms with special adaptive strategies
to the extreme environmental conditions (Ordoñez et al.
2009). Former studies focused mainly on other lakes of the
Dry-Andes (Demergasso et al. 2008; Dorador et al. 2008,
2013; Cabrol et al. 2009; Farías et al. 2009, 2014; Lynch
Extremophiles (2016) 20:603–620
et al. 2012; Scott et al. 2015), and research was also performed on similar environments, for example Antarctic
lakes and permafrost (Goordial et al. 2016; Aislabie et al.
2013; Antibus et al. 2012; Cowan et al. 2002) and high-altitude lakes or permafrost of the Tibetan Plateau (Wu et al.
2006; Xing et al. 2009; Zhang et al. 2007). Understanding
the microbial processes in degrading permafrost is also an
important issue, since thawing is proven to cause shifts in
microbial communities, functional gene abundances which
might be responsible for changes in global carbon cycle
(Mackelprang et al. 2011; Tas et al. 2014; Hultman et al.
2015).
The aim of the present work was to reveal the hitherto
unknown bacterial communities inhabiting the sediments
of five different high-altitude lakes located in the Ojos del
Salado volcano, Dry-Andes. To gain information about the
bacterial diversity and compare the community structures,
16S rRNA gene based denaturing gradient gel electrophoresis and molecular cloning methods were applied.
Materials and methods
Description of sampling sites
The studied five high-altitude lakes are located near the
border between Chile and Argentina in the area of Ojos del
Salado volcano. Geographical location and GPS coordinates
of the lakes are indicated in Fig. 1 and Table 1. Laguna Santa
Rosa and Laguna Verde are permanent saline lakes located at
3770 and 4350 m.a.s.l., respectively. Because of the intense
evaporation (1000 mm year−1) and very small amount of
precipitation (170 mm year−1), the surface area of Laguna
Santa Rosa and Laguna Verde has been constantly shrinking
(Hiner 2009). This phenomenon caused an increase in the
salt concentration of the lake water and accumulation of salt
Fig. 1 Geographic location of Ojos del Salado (red star) in the map of South-America (a) and the sampling sites (yellow crosses) (b)
13
Extremophiles (2016) 20:603–620
605
Table 1 Geographic parameters of the lakes in the area of Ojos del Salado volcano, location and physical characteristics of the sampling sites,
numerical data of DGGE and 16S rRNA gene clone library analysis
3770
Laguna
Verde
4350
2
1.2
Laguna Santa Rosa
Elevation (m.a.s.l.)
Lake area (km2, estim.)
Maximum depth (m, estim.)
Sample code
GPS coordinates
Water temperature (°C)
Air temperature (°C)
Lake I
Lake II
5900
5900
Lake III
6500
15
0.0025
0.0025
0.006
4
1
1
1
AS-12
AS-13
AS-14
AS-5
AS-6
AS-9
AS-10
AS-1
AS-2
AS-3
AS-4
27.081085 S 27.081085 S 26.890288 S 27.088495 S 27.088495 S 27.088360 S 27.088396 S 27.109925 S 27.110914 S 27.110900 S 27.110206 S
69.175022 69.175022 68.486858 68.534330 68.534330 68.531410 68.531532 68.550184 68.550108 68.550121 68.550885
7
7
36
7
7
7
7
0
36
36
2
5
5
7
0
0
0
0
-5
-5
-5
-5
9.5
9.5
8.8
6.6
6.6
6.8
6.8
4.8
2.1
2.1
4.8
6540
6540
6260
1230
1230
1081
1542
n.e.
1360
n.e.
n.e.
0-5
10-15
0-5
0-5
10-15
20
0-5
0-5
0-5
0-5
0-5
Number of DGGE bands
14
30
22
27
32
26
26
13
5
9
15
Number of ARDRA groups
n.e.
43
42
27
n.e.
n.e.
42
28
25
n.e.
n.e.
Number of identified molecular clones
n.e.
38
34
21
n.e.
n.e.
33
27
24
n.e.
n.e.
Clones with >97% similarity to a described species
n.e.
4
1
7
n.e.
n.e.
19
1
0
n.e.
n.e.
Clones with >97% similarity to a molecular clone
n.e.
13
2
7
n.e.
n.e.
4
9
7
n.e.
n.e.
pH
Conductivity (μS cm-1)
Sediment sampling depth (cm)
ne not examined, estim. estimated
crystals on the lakeside. Laguna Verde is fed by several cold
and warm (34–36 °C) springs. Two seasonal nameless lakes
(Lake I and II) found at 5900 m.a.s.l., are derived from the
melting permafrost, and both can be characterized by shallow clear water bodies. These lakes were partially ice covered at the time of sampling. The third seasonal lake (Lake
III) is located at 6500 m.a.s.l. This lake is surrounded by several fumaroles and fed by meltwater of perennial snowfields
(1 °C) and a small warm (36 °C) creek. At the time of sampling this lake was completely frozen down to the bottom.
Presence of warm springs near Laguna Verde and the warm
spring and fumaroles at the 6500 m elevation site are signs of
volcanic activity.
Sampling was accomplished during the Földgömb Atacama Climate Monitoring Expedition in February 2014.
Samples were collected into 50 ml sterile Falcon tubes
from different sediment layers of all five lakes as indicated
in Table 1, and were stored between −5 and 10 °C until
being processed in the laboratory.
DNA extraction and PCR amplification
The community DNA was isolated using the Ultra Clean
Soil Kit (MO Bio Inc., CA, USA) according to the manufacturer’s instructions. The 16S rRNA gene was amplified
by PCR using Bacteria-specific 27 forward (5′-AGAGTTT
GATCMTGGCTCAG-3′) (Lane 1991) and 1401 reverse
(5′-CGGTGTGTACAAGACCC-3′) (Nübel et al. 1996)
primers. The following temperature protocol was used
for bacterial PCR: initial denaturation at 95 °C for 5 min,
followed by 32 cycles of denaturation at 94 °C for 30 s,
annealing at 52 °C for 45 s and elongation at 72 °C for
60 s, and a final extension at 72 °C for 10 min. The PCR
reaction mixture contained 200 µM of each deoxynucleoside triphosphate, 1 U of LC Taq DNA Polymerase
(recombinant) (Fermentas, Lithuania), 1X Taq buffer with
(NH4)2SO4 (Fermentas, Lithuania), 2 mM MgCl2, 0.65 mM
of each primer, and approximately 20 ng of genomic DNA
template in a total volume of 25 µL. Both community DNA
isolates and PCR products were visualized in 1 % agarose
gel stained with ECO Safe Nucleic Acid Staining Solution
(Avegene, Taiwan) using UV excitation.
Denaturing gradient gel electrophoresis (DGGE)
A semi-nested PCR was carried out on the 16S rRNA gene
amplicons using a GC-clamp (5′-GC clamp-CAGAGTTTGATCCTGGCTCAG-3′) containing 27 forward and 519
reverse (5′-GTNTTACNGCGGCKGCTG-3′) primers
(Muyzer et al. 1993). The reaction mixture and temperature
protocol were the same as mentioned above. Electrophoresis of PCR products was run for 14 h using an INGENY
phorU-2 electrophoresis system in a 7 % polyacrylamide
gel containing 40–60 % gradient of denaturants (40 %
formamide and 7 M urea is defined as 100 %) in 1X TAE
buffer at 100 V and 60 °C. Following the run, the gel was
stained with ethidium-bromide, and detected under UV
light.
Discrete bands were excised from the gel with a sterile
scalpel, and after an overnight incubation in 30 μl of distilled water, the extracted DNA was amplified with the 27
forward and 519 reverse primers using the same reaction
mixture and temperature profile as described previously.
The DGGE profiles of the samples were compared, and
an UPGMA (unweighted pair group method using arithmetic mean algorithm) dendrogram (Smalla et al. 2007) was
constructed based on the position and presence of bands in
the different lanes (representing the samples) using Total
Lab (TL 120) version 2006 (Nonlinear Dynamics Inc.,
Newcastle upon Tyne, UK) software.
13
606
Construction of molecular clone libraries
The purified (EZ-10 Spin Column PCR Purification Kit,
Bio Basic, Canada) 27 forward—1401 reverse PCR products were ligated into a TA-cloning vector (pGEM-T Vector System, Promega, WI, USA) according to the manufacturer’s instructions, and transformed into competent E.
coli JM109 cells. For blue/white selection, the transformed
cells were spread on LB agar plates containing 100 µg ml−1
ampicillin, 80 µg ml−1 X-Gal and 0.5 mM IPTG, and incubated overnight at 37 °C. White colonies were picked for
the clone libraries, and DNA was extracted from the E. coli
cells by incubating the cultures at 98 °C for 5 min, and pelleting the cell fragments by centrifugation with 4500 rcf for
5 min. Insert sequences of each clone were amplified with
standard M13 forward (5′-GTAAAACGACGGCCAGT-3′)
and M13 reverse (5′-CAGGAAACAGCTATG-3′) primers
(Messing 1983) followed by a nested PCR with the original
27 forward and 1401 reverse primers. The thermal profiles
of PCRs were the same as described previously.
Extremophiles (2016) 20:603–620
forward primers using the automated Sanger-method by LGC
Ltd. (Berlin, Germany). The quality of each chromatogram
was checked with the help of the Chromas software, and low
quality ends were trimmed (Technelysium Pty Ltd., Australia). Taxonomic relationships of the partial 16S rRNA gene
sequences were determined by EzTaxon-e database (Kim et al.
2012). The 16S rRNA gene sequences were deposited into
GenBank under accession numbers LN929548-LN929732.
In the case of molecular clones, phylogenetic relations
of sequences at least 97 % similarity to an uncultured environmental clone or described species (Stackebrandt and
Goebel 1994; Vandamme et al. 1996; Tindall et al. 2010)
found in GenBank were analyzed by construction of neighbor joining trees with MEGA6 software (Tamura et al.
2013). Sequences (along with the nearest cultivated relative or molecular clone sequences derived from GenBank)
were aligned by ClustalW (Larkin et al. 2007), the number
of bootstrap replications was 1000 and Kimura 2-parameter
(Kimura 1980) model was applied.
ARDRA grouping and rarefaction analysis
Results and discussion
PCR amplicons were grouped on the basis of their amplified ribosomal DNA restriction analysis (ARDRA) patterns
produced with restriction enzymes as described by MassolDeya et al. (1995). The applied restriction enzymes (Hin6I
and BsuRI, Fermentas, Lithuania) were selected based on
their different specificity (3′-G^CGC-5′ and 3′-GG^CC-5′,
respectively) that resulted in different restriction pattern of
PCR products allowing us to make ARDRA groups (Krett
et al. 2013; Borsodi et al. 2012). Each reaction mixture
contained 2 µl R/Tango Buffer (Fermentas), 7.76 µl sterile distilled water, 0.24 µl of enzyme and 10 µl of PCR
product. Digestions were made at 37 °C, for 3 h. Digestion products were separated in 2 % agarose gel on 80 V
for 80 min, stained with ECO Safe Nucleic Acid Staining
Solution (Avegene, Taiwan), and visualized by UV excitation. Clones were grouped according to their different
ARDRA patterns. At least one representative from each
ARDRA group was sequenced.
Rarefaction analysis was carried out to check, whether
the number of clones analyzed was satisfactory to estimate
diversity within the clone libraries and to compare the diversity of clone libraries. Rarefaction curves were produced
with the software Past3 (Hammer et al. 2001) by plotting
the number of 16S rRNA gene clones in each clone libraries
against the number of clones in different ARDRA groups.
Denaturing gradient gel electrophoresis
Sequencing, identification and phylogenetic analysis
PCR products from the excised DGGE bands and representative clones of each ARDRA group were sequenced with 27
13
To compare the bacterial community structures of the sediment samples from the lakes located in the Ojos del Salado
region, similarity dendrogram was constructed on the basis
of the DGGE patterns (Fig. 2). Three groups were formed
according to the altitude of the studied lakes. Samples from
the highest altitude were clearly separated from the others;
furthermore, they formed pairs (AS-2 and AS-3; AS-1 and
AS-4) in relation to the different temperatures of the sampling
sites. Samples from Lake I (AS-5, AS-6) and Lake II (AS9, AS-10) located at 5900 m.a.s.l. formed the second group.
The two sediment samples (AS-12 and AS-13) from Laguna
Santa Rosa (3770 m.a.s.l.) formed the third group together
with the sample (AS-14) from Laguna Verde (4350 m.a.s.l.).
It is interesting to note that although sample AS-14 derived
from a warm environment, it did not contain typical common
bands with the other two warm samples (AS-2 and AS-3).
Altogether 85 different bands were detected from the gel.
Number of DGGE bands for each sample is indicated in
Table 1. The smallest amount of bands (thus the least diverse
community) was revealed in samples derived from 6500 m.a.s.l.
(AS-1, AS-2, AS-3, AS-4), especially in the sediments of the
warm shallow creek (AS-2, AS-3). Number of bands from samples at 3770 and 5900 m.a.s.l. consisted of 14–32 bands without
a clear relationship to the altitude of the lakes or the depth of
sediment. At the same time, samples from lakes I and II found at
5900 m.a.s.l. contained surprisingly large amount of bands (26–
32) suggesting high genetic diversity. In addition, the difference
between the band numbers of samples AS-12 (30) and AS-13
Extremophiles (2016) 20:603–620
607
Fig. 2 Similarity UPGMA
dendrogram of the DGGE
molecular fingerprints of the
sediment samples from lakes
of Ojos del Salado (elevation is
indicated after the code of the
sample. The excised bands are
marked with arrows)
Table 2 Phylogenetic affiliation of 16S rRNA gene sequences of the excised DGGE bands according to EzTaxon
DGGE band Sample code Accession number Similarity (%) E value
Coverage (%) Closest taxon in EzTaxon (accession number)
Firmicutes
An_9
100
LN929727
97 (179/184)
7.00E−75
Gemmatimonadetes
An_3
AS-6
AS-12
LN929726
85 (384/450)
8.00E−113 100
Gemmatimonas aurantiaca (AP009153 )
Proteobacteria
An_29
AS-1
Rhodanobacter umsongensis (FJ821731)
LN929732
98 (461/469)
0
An_1
AS-4
LN929724
91 (418/461)
2.00E−169 100
Thiobacillus thiophilus (EU685841)
An_14
AS-13
LN929731
98 (461/471)
0
100
Thiobacillus thiophilus (EU685841)
An_2
AS-6
LN929725
98 (429/436)
0
100
Noviherbaspirillum psychrotolerans (JN390675)
An_13
AS-9
LN929730
99 (463/470)
0
94
Noviherbaspirillum psychrotolerans (JN390675)
An_10
AS-10
LN929728
97 (398/412)
4.00E−175 100
Brevundimonas faecalis (FR775448)
An_11
AS-10
LN929729
98 (460/471)
0
Thermomonas brevis (AJ519989)
(14) was more than twice, despite the fact that both originated
from Laguna Santa Rosa located at 3770 m.a.s.l.
From the DGGE gel a total of 14 discrete bands were
excised and sequenced out of which 9 were successfully
identified. They belonged to three different phyla (Firmicutes, Gemmatimonadetes and Proteobacteria) as it is indicated in Table 2. At species level, DGGE sequences were
closely related to e.g., the extreme halophilic Halanaerobium
sehlinense, firstly isolated from the Tunisian hypersaline lake
Sehline Sebkha (Abdeljabbar et al. 2013) and the psychrotolerant Noviherbaspirillum psychrotolerans, described from
Larsemann Hills, Antarctica (Bajerski et al. 2013).
16S rRNA gene based molecular clone libraries
Based on the DGGE similarity dendrogram, 6 sediment
samples (at least one from each altitude) were chosen for
100
Halanaerobium sehlinense (JN381500)
100
a detailed phylogenetic analysis by molecular clone libraries. Both samples AS-1 and AS-2 were derived from the
upper 0–5 cm sediment layer at the highest altitude sampling site (Lake III), from a cold (0 °C) and a warm (36 °C)
environment, respectively. From the lakes at 5900 m.a.s.l.
two samples were chosen. Both samples AS-5 (Lake I) and
AS-10 (Lake II) were taken from 0 to 5 cm sediment depth
but on the DGGE similarity dendrogram they clearly separated from each other. Clone libraries from Laguna Verde
(sample AS-14) and Laguna Santa Rosa (sample AS-13)
were also examined in detail.
Following the ARDRA grouping, 177 representatives
were sequenced and identified from the altogether 431
molecular clones. The rarefaction curves (Fig. 3) did not
reach asymptotes, indicating that further diversity would
have been revealed by the analysis of a higher number of
clones. The presence of the most diverse communities was
found in the saline lakes at lower altitudes (samples AS-13,
13
608
AS-14) while the highest elevation sites (samples AS-1,
AS-2) hosted the least diverse bacterial assemblages on the
basis of rarefaction curves.
In the six samples, members of 17 different phyla or
candidate divisions (Aquificae, Deinococcus-Thermus,
Chloroflexi, Nitrospirae, Cyanobacteria, Chlorobi, Proteobacteria, Firmicutes, Actinobacteria, Acidobacteria, Bacteroidetes, Verrucomicrobia, Caldithrix, Gemmatimonadetes,
TM7, OD1, WS5) were detected. However, major differences were observed in the composition of the bacterial
assemblages from different sampling sites. The relative
abundance of molecular clones among the bacterial phyla
per sample is shown in Fig. 4.
Members of Proteobacteria were present in each clone
library, and in 4 cases (AS-1, AS-2, AS-5, AS-13) Betaproteobacteria was the most abundant. The AS-10 clone library
was predominated with Gammaproteobacteria. In AS-14,
unlike other clone libraries, the class Betaproteobacteria
was not detected but the members of Epsilonproteobacteria
were present. Representatives of Acidobacteria, Bacteroidetes, Actinobacteria and Cyanobacteria were also detected
at least in three clone libraries.
The clone library constructed from the sediment sample of the extreme saline Laguna Santa Rosa (AS-13) contained 10 phyla, and showed the most diverse community
structure amongst the examined lakes. Proteobacteria was
the most abundant phylum with four classes (Alpha-, Beta-,
Gamma-, and Deltaproteobacteria), but members of Acidobacteria, Bacteroidetes, Actinobacteria, Chloroflexi and
Verrucomicrobia were also abundant (Fig. 4). In this clone
library altogether 37 ARDRA groups were separated. However, most of the molecular clones showed less than 97 %
similarity to any described species. Only a few molecular
clones (Fig. 5) could be identified at species level (with
higher than 97 % of 16S rRNA gene sequence similarity
to described species) (Stackebrandt and Goebel 1994; Vandamme et al. 1996; Tindall et al. 2010). They were closely
related to the species Rhodoferax ferrireducens (Betaproteobacteria), the first described non-phototrophic member
of the genus (Finneran et al. 2003) which grows via dissimilatory Fe(III) reduction, and R. antarcticus (Madigan
et al. 2000), a moderately psychrophilic (growth optimum
15–18 °C) purple non-sulfur bacterium first isolated from
Antarctic microbial mats. Representatives of the species
Gillisia hiemivivida (Bowman and Nichols 2005) were present in this clone library, too. It is a psychrophilic and halophilic bacterium. Sequences closely related to the chemolithotrophic nitrite-oxidizing species of Nitrospira marina
(Watson et al. 1986) was also detected. Sequences showing
96 % similarity to species Herminiimonas glaciei (Loveland-Curtze et al. 2009) detected in this clone library represented one of the dominant DGGE bands in the sample
from Lake II found at 5900 m.a.s.l. It is worth mentioning
13
Extremophiles (2016) 20:603–620
Fig. 3 Rarefaction curves for the different ARDRA patterns of 16S
rRNA gene clones from lake sediment samples of the lakes located in
the area Ojos del Salado volcano
Fig. 4 Distribution of members of different phyla in the 16S rRNA
gene clone libraries constructed from sediment samples of the lakes
of Ojos del Salado, Dry-Andes
that the largest ARDRA group of this sample showed the
closest similarity to an uncultured environmental clone
(Par-s-7, EF632909) from the high-altitude Chilean Altiplano (Dorador et al. unpublished data).
Fig. 5 Neighbor-joining phylogenetic tree based on the 16S rRNA ▸
gene sequence data of bacterial clones from sediment samples of
the lakes of Ojos del Salado showing >97 % similarity to described
species. The number of members of the ARDRA groups is indicated
after the representative clone
Extremophiles (2016) 20:603–620
609
AS-10-79 (LN929646) / 2 clones
AS-5-51 (LN929610) / 2 clones
AS-10-47 (LN929635) / 1 clone
55 71 Caenimonas terrae strain SGM1-15 (GU181268) / Soil (Korea)
99 AS-10-55 (LN929638) / 7 clones
AS-5-3 (LN929598) / 2 clones
45 98
Acidovorax delafieldii (AF078764) / n.d.
94 Acidovorax facilis strain THWCSN37 (GQ284428) / Natural spring sediment (India)
63 A S-13-23 (LN929661) / 2 clones
70
Betaproteobacteria
AS-13-3 (LN929653) / 3 clones
100
Rhodoferax ferrireducens (AF435948) / Subsurface sediment, Oyster Bay (USA)
100
AS-5-79 (LN929615) / 1 clone
100 Polaromonas cryoconiti strain Cr4-35 (HM583567) / Pitztaler Jöchl glacier cryoconite (Austria)
AS-10-59 (LN929640) / 1 clone
100
99 Hydrogenophaga palleronii (AF019073) / n.d.
86 AS-10-76 (LN929645) / 1 clone
99
AS-5-20 (LN929602) / 3 clones
AS-10-30 (LN929628) / 1 clone
100
Janthinobacterium svalbardensis strain JA-1 (DQ355146) / Austre Broggerbreen glacier (Spitzbergen)
100
AS-1-71 (LN929568) / 1 clone
Rhodanobacter umsongensis strain GR24-2 (FJ821731) / Soil of a ginseng field (China)
100
AS-10-83 (LN929647) / 3 clones
99
Pseudoxanthomonas wuyuanensis (JN247803) / Saline-alkali soil (China)
Gammaproteobacteria
99 AS-10-16 (LN929625) / 7 clones
93
Thermomonas brevis GZ436 (KF923805) / Soil of a coal mine (China)
71
AS-10-34 (LN929631) / 2 clones
95 Lysobacter koreensis (AB166878) / Soil of a ginseng field (Korea)
AS-10-95 (LN929651) / 2 clones
Gammaproteobacteria
100 Cellvibrio fibrivorans R4079 (AJ289164) / Soil (Belgium)
100 AS-10-6 (LN929620) / 4 clones
Brevundimonas variabilis strain ATCC 15255 (AJ227783) / n.d.
100 AS-10-4 (LN929619) / 1 clone
100
Devosia glacialis strain Cr4-44 (HM474794) / Pitztaler Jöchl glacier cryoconite (Austria)
Alphaproteobacteria
100 AS-10-42 (LN929633) / 4 clones
62
Catellibacterium aquatile strain A1-9 (EU313813) / Daqing reservoir freshwater (China)
100
AS-14-7 (LN929695) / 1 clone
100
Roseinatronobacter monicus strain ROS 35 (DQ659236) / Mono Lake (USA)
AS-13-82 (LN929687) / 1 clone
Nitrospira
100
Nitrospira marina (X82559) / Heating system (Russia)
100 AS-10-31 (LN929629) / 10 clones
Microcoleus vaginatus PBP-D-KK1 (EF654072) / n.d.
Cyanobacteria
100
AS-10-60 (LN929641) / 5 clones
100 Crinalium epipsammum SAG 22.89 (AB115964) / n.d.
100 AS-10-48 (LN929636) / 1 clone
Hymenobacter roseosalivarius strain AA718 (Y18833) / Soil (Antarctica)
99 AS-5-56 (LN929612) / 1 clone
100
AS-10-9 (LN929622) / 1 clone
Pedobacter daechungensis strain Dae 13 (AB267722) / Soil of a ginseng field (China)
Bacteroidetes
100 AS-13-18 (LN929658) / 2 clones
94
Gillisia hiemivivida strain IC154 (AY694006) / Maritime habitat (Antarctica)
99 AS-5-29 (LN929604) / 10 clones
100
Flavobacterium glaciei strain 0499 (DQ515962) / Glacier (China)
100
AS-5-80 (LN929616) / 7 clones
92
Flavobacterium psychrophilum strain IFO 15942 (AB078060) / n.d.
100 AS-10-10 (LN929623) / 1 clone
Actinobacteria
Nocardioides alpinus strain Cr7-14 (GU784866) / Alpine glacier cryoconite (Austria)
AS-10-11 (LN929624) / 1 clone
Deinococcus-Thermus
100 Deinococcus marmoris strain AA-63 (JNIV01000230) / Marble rock (Antarctica)
96
79
63
93
61
68
41
100
78
0.05
AS-1: 6500 m (cold water)
AS-2: 6500 m (warm water)
AS-5: 5900 m (cold water)
AS-10: 5900 m (cold water)
AS-14: 4350 m (warm water)
AS-13: 3770 m (cold water)
13
610
The clone library constructed from the sediment sample
of the extreme saline Laguna Verde (AS-14) was unique
among the studied samples. Members of altogether 6 phyla
were present in this clone library (Fig. 4) including four
classes of Proteobacteria (Alpha, Gamma- Delta- and Epsilonproteobacteria). Most of the identified molecular clones
showed less than 97 % similarity to any environmental
clones or formerly described species of bacteria (Figs. 5,
6). The most abundant groups showed only 90–91 % similarity to environmental clones from the phyla Caldithrix
and Chloroflexi, respectively. Only one minor representative could be identified as member of species Roseinatronobacter monicus (Boldareva et al. 2007), an alkaliphilic halotolerant Alphaproteobacteria described from the
hypersaline Mono Lake, California. Other ARDRA group
representatives showed the highest similarity to uncultured clones from hot environment of methane-rich mud
volcanoes, anaerobic hydrothermal vents or hot-springs in
Greenland, Columbia or the Philippines. The relatively low
sequence similarity rates suggest high abundance of undescribed species within this sample.
Bacterial communities hosted by two permafrost derived
lakes found at 5900 m.a.s.l. were very similar according to
the clone libraries AS-5 and AS-10 (Fig. 4). The two clone
libraries consisted of members of 7 phyla and shared many
similar sequences, and both of them included molecular
clones closely related to ones described from cryoconites,
glaciers, permafrost sediments or freshwater lakes (Fig. 6).
Among the molecular clone representatives, species
Janthinobacterium svalbardensis (Avguštin et al. 2013)
was detected in both samples. This psychrotrophic (growth
occurs at 2–25 °C) bacterium was first described from a
glacier in the Spitsbergen. Members of the genus Flavobacterium were also present in both clone libraries, and
formed the most abundant ARDRA group in sample AS-5.
Representatives of cold-adapted species F. glaciei (Zhang
et al. 2006) and F. psychrophilum (Nakagawa et al. 2002)
were detected in the sediment of Lake I (AS-5). Similarly, species F. psychrolimnae (Van Trappen et al. 2005)
was described from microbial mats in the Antarctic Lake
Fryxell, representatives of which were exposed from sediment sample AS-10 of Lake II. Betaproteobacteria was the
second most abundant group in AS-5, and the presence of
species Polaromonas cryoconitii was detected. This psychrophilic bacterium was first isolated from a cryoconite of
Pasterze/Großglockner glacier in the Hohe Tauern, Austria
(Margesin et al. 2012).
Interestingly, in sample AS-10 members of phylum
Cyanobacteria were also detected, and according to the
13
Extremophiles (2016) 20:603–620
Fig. 6 Neighbor-joining phylogenetic tree based on the 16S rRNA ▸
gene sequence data of bacterial clones from sediment samples of the
lakes of Ojos del Salado showing >97 % similarity to an uncultured
bacterial clone. The number of members of the ARDRA groups is
indicated after the representative clone
number of molecular clones they were dominant, however,
this phylum was absent in sample AS-5 (Fig. 4).
A minor ARDRA group in sample AS-10 was closely
related to Hymenobacter roseosalivarius, a species of
phylum Bacteroidetes (Fig. 5). This strictly aerobic, oligotrophic bacterium was isolated from soil samples of the
Antarctic polar desert, Dry Valleys (Hirsch et al. 1998).
In sample AS-10, only one molecular clone was related
to Actinobacteria. It showed 97 % similarity to species
Nocardioides alpinus, an actinomycete isolated from a
cryoconite of Pitztaler Jöchl glacier in the Ötztaler Alps in
Tyrol, Austria (Zhang et al. 2012).
AS-10 was the only sample where presence of Chlorobi
was observed (Fig. 4). A molecular clone showed high similarity to an unknown environmental sequence from moss
pillars in a freshwater lake, Hotoke Ike, Antarctica (Nakai
et al. 2012).
The only member of the phylum Deinococcus-Thermus
in this study was detected in the sample AS-5 (Fig. 4).
Deinococcus marmoris is, a desiccation and UV radiation
tolerating bacterium was described from Antarctic marble
(Hirsch et al. 2004). Molecular clones related to the members of Gemmatimonadetes were also present in the sample
AS-5 (Fig. 4). These clones showed the highest similarity
to other environmental sequences, for example an environmental clone derived from Mendenhall glacier, USA (Sattin et al. 2009) and another from Socompa volcano in the
Dry-Andes (Costello et al. 2009).
Two sediment samples (AS-1 and AS-2) from Lake
III found at 6500 m.a.s.l derived from relative proximity,
however, have different physical and chemical characteristics (Table 1). AS-1 was a sediment sample from the shore
affected by melt water of the ice body of the lake. AS-2
was a sediment sample under the warm and extremely
acidic spring feeding the lake. Both DGGE and molecular clone library results showed characteristic differences
between the bacterial communities inhabiting these sediment samples.
In the case of sample AS-1, dominance of Acidobacteria was revealed (Fig. 4). The most abundant ARDRA
group was closely related to Granulicella arctica (Männistö et al. 2012). Occurrence of members of genus Granulicella is frequent in Polar Regions, several species were
Extremophiles (2016) 20:603–620
611
AS-2-10 (LN929580) / 7 clones
Uncultured bacterium clone A1 e2 (HQ317061) / Acid rock drainage (Antarctica)
AS-1-86 (LN929574) / 2 clones
43
100 Uncultured bacterium clone V6 44 (JF267704) / Coal mine heap, Sokolov (Czech Republic)
AS-13-30 (LN929664) / 2 clones
86
100
Uncultured bacterium clone 4.4.9 (EU528206) / Kentucky lake sediment (USA)
Betaproteobacteria
100
AS-5-55 (LN929611) / 5 clones
Uncultured bacterium clone GA091 (EU636041) / Collins glacier (Antarctica)
86
AS-10-7 (LN929621) / 3 clones
100 Uncultured bacterium clone 1250 (FM209333) / Sand soil, Negev desert (Israel)
AS-5-45 (LN929607) / 1 clone
100
Uncultured bacterium clone reservoir-108 (JF697489) / Stream of Dianchi lake (China)
98 AS-13-13 (LN929656) / 2 clones
100
AS-13-75 (LN929685) / 1 clone
Uncultured bacterium clone Alchichica AL52 2 1B 85 (JN825498) / Alchichica alkaline lake (Mexico)
AS-10-71 (LN929643) / 4 clones
100 Uncultured bacterium clone SR-O-03-32 (AM905315) / High-Arctic intertidal beach sediment (Norway) Gammaproteobacteria
AS-10-84 (LN929648) / 1 clone
100
Uncultured bacterium clone KD2-14 (AY218573) / Penguin dropping sediment, Ardley Island (Antarctica)
AS-13-36 (LN929667) / 1 clone
100
Uncultured bacterium clone FFCH8865 (EU134765) / Soil of mixed grass prairie (USA)
AS-13-46 (LN929674) / 1 clone
100
AS-13-19 (LN929659) / 1 clone
94
Uncultured bacterium clone TX1A 146 (FJ152698) / Alkaline saline soil, former Lake Texcoco (Mexico) Alphaproteobacteria
AS-1-14 (LN929555) / 1 clone
100
Uncultured bacterium clone JFJ-ICE-Bact-04 (AJ867748) / Melted-ice water (Switzerland)
AS-2-5 (LN929578) / 10 clones
100
Firmicutes
AS-2-3 (LN929577) / 15 clones
91
Uncultured bacterium clone Central-Bottom-cDNA clone89 (HE604029) / Acidic lignite mine lake (Germany)
100 AS-14-91 (LN929721) / 5 clones
Uncultured bacterium clone LEGE 07319 (HM217045) / Estuary (Portugal)
Cyanobacteria
AS-14-6 (LN929694) / 1 clone
100
Uncultured bacterium clone SINH475 (HM128037) / Xiaochaidan lake (Tibet)
100
AS-13-63 (LN929678) / 1 clone
Chloroflexi
Uncultured bacterium clone FCPT687 (EF515885) / Grassland soil (CA USA)
AS-1-74 (LN929570) / 3 clones
97 AS-1-44 (LN929562) / 6 clones
TM7
100
AS-1-63 (LN929567) / 1 clone
82
Uncultured bacterium clone C129 (AF507687) / pinyon-juniper forest soil, Arizona (USA)
100
AS-5-31 (LN929605) / 1 clone
Uncultured bacterium clone P3 (DQ351728) / Soil of LaGorce mountains (Antarctica)
Bacteroidetes
AS-13-60 (LN929676) / 1 clone
100 Uncultured bacterium clone 20m-41 (GU061294) / Intertidal beach seawater, Yellow Sea (Korea)
100 AS-5-9 (LN929599) / 2 clones
Uncultured bacterium clone 5B 04E (JX098559) / 6000 meters elevation mineral soils (Atacama desert)
100
99 AS-5-61 (LN929613) / 2 clones
Uncultured bacterium clone AK4AB2 04A (GQ396928) / Recently deglaciated soil, Mendenhall glacier (USA) Gemmatimonadetes
99
AS-13-72 (LN929682) / 1 clone
100 Uncultured bacterium clone KC-4 (EU421850) / Soil under a glacier, Lahaul-spiti Valley (Indian Himalaya)
AS-13-5 (LN929655) / 1 clone
100 Uncultured bacterium clone BN23 (HQ190281) / Zhongynan oil field (China)
100
AS-1-80 (LN929573) / 2 clones
Uncultured bacterium clone P23 (DQ351736) / Soil of LaGorce mountains (Antarctica)
100 AS-2-12 (LN929582) / 17 clones
Actinobacteria
Uncultured bacterium clone OY07-C082 (AB552368) / Volcanic ash deposit (Japan)
AS-1-32 (LN929560) / 1 clone
100
Uncultured bacterium clone OY07-C186 (AB552456) / Volcanic ash deposit (Japan)
71 AS-5-92 (LN929618) / 2 clones
100
Uncultured bacterium clone A01 SB2A (FJ592652) / High-altitude warm fumarole soil, Socompa volcano (Andes) Verrucomicrobia
AS-13-73 (LN929683) / 4 clones
92 Uncultured bacterium clone Par-s-7 (EF632909) / High-altitude freshwater sediment (Chilean Altiplano)
99 AS-13-41 (LN929671) / 1 clone
84
Uncultured bacterium clone RB41(Z95722) / n.d.
AS-13-77 (LN929686) / 2 clones
100
99 Uncultured bacterium clone GTM2314fO9 (JN615809) / Microbial mat from lava cave (Portugal)
AS-10-44 (LN929634) / 2 clones
100 AS-5-50 ( LN929609) / 1 clone
61
Uncultured bacterium clone EME017 (EF127595) / Ice of Dry Valleys (Antarctica)
Acidobacteria
87 AS-2-56 (LN929590) / 1 clone
100 AS-2-85 (LN929596) / 1 clone
AS-2-81 (LN929594) / 1 clone
65
89 Uncultured bacterium clone RT8-ant02-d03-W (JF807641) / Rio Tinto water (Spain)
AS-1-77 (LN929571) / 3 clones
Uncultured bacterium clone B146 (FJ466009) / Kilauea volcanic deposit (Hawaii)
100 90
AS-1-29 (LN929558) / 1 clone
95
Uncultured bacterium clone ERF-1A1 (DQ906050) / Rhizosphere on the bank of Rio Tinto (Spain)
99
99
98
84
54
45
96
58
16
76
100
17
36
54
99
74
24
89
99
22
51
72
0.02
AS-1: 6500 m (cold water)
AS-2: 6500 m (warm water)
AS-5: 5900 m (cold water)
AS-10: 5900 m (cold water)
AS-14: 4350 m (warm water)
AS-13: 3770 m (cold water)
13
612
described from tundra soils in Finland. Species G. arctica was reported to be dominant in acidic soils, showing
growth at only 4 °C, and being tolerant to several freeze–
thaw cycles (Männistö et al. 2012). The second most
abundant major ARDRA group was related to environmental clones from forest soils belonging to the candidate phylum TM7 (Dunbar et al. 2002). In AS-1 members
of Actinobacteria were only far relatives of three species,
Glaciihabitans tibetensis, isolated from glacial meltwater
affected sediment in Tibet (Li et al. 2014), the psychrophilic (optimum at 1–15 °C) Alpinimonas psychrophila,
described from cryoconites of Rettenbach glacier, Austria (Schumann et al. 2012), and the halotolerant (<4 %
NaCl) Calidifontibacter indicus, described from a hot
spring in India (Ruckmani et al. 2011). Other molecular
clones from this library belonging to the phylum Actinobacteria showed the highest sequence similarity to uncultured bacteria detected in soil samples from La Gorce
mountains, Antarctica (Aislabie et al. 2006) and from
volcanic sediments in Japan (Fujimura et al. 2012).
In sample AS-1, representatives of three classes of Proteobacteria (Alpha-, Beta-, and Gammaproteobacteria)
were detected (Fig. 4). Members of Alphaproteobacteria were related to molecular clones from mineral soil at
6000 m.a.s.l, in the Dry-Andes (Lynch et al. 2012). A
molecular clone was closely related to the mesophilic species Rhodanobacter umsongensis (Kim et al. 2013) which
was also detected by DGGE analysis. Some molecular
clones of the sample AS-1 showed low similarity to molecular clones from soil of Princess Elisabeth Station, Antarctica (Peeters et al. 2011) and to species Mucilaginibacter
frigoritolerans a freeze-thaw cycle tolerating bacterium
(Männistö et al. 2010) within phylum Bacteroidetes.
In sample AS-2 more than 33 % of the clones showed
the highest similarity to environmental sequences belonging to the phylum Firmicutes (Fig. 4). This phylum was
also present in the other warm sample AS-14. Phylum Actinobacteria was the second most abundant group. Molecular clones showed the highest similarity to environmental
sequences, e.g., from the formerly mentioned volcanic sediment in Japan (Fujimura et al. 2012), the Mendenhall glacier in the United States, and clones from the acidic river
Rio Tinto, Spain (García-Moyano et al. 2012). Molecular
clones of sample AS-2 belonging to Acidobacteria were
also similar to sequences from these acidic environments.
Representatives of the phylum Cyanobacteria were also
detected similar to the other warm-water sample AS-14.
13
Extremophiles (2016) 20:603–620
Comparison of the communities inhabiting the lake
sediments of Ojos del Salado and similar volcanic,
high‑altitude lake or permafrost sediments
Communities detected in the lake sediments of Ojos del
Salado consisted of sequences related mainly to psychrophilic, acidophilic or halophilic microorganisms. The
composition of the bacterial communities was similar to
those reported from volcanic, high-altitude or permafrost
sediments (Hultman et al. 2015; Taş et al. 2014). In these
areas low organic material in soils, freeze-thaw cycles and
aridity are probably the main components of the environmental stress which resulted in similar bacterial community composition at phylum level. Proteobacteria, Actinobacteria and Bacteroidetes were the most frequently
detected phyla in almost all studied environments including the samples from the area of Ojos del Salado, as well
(Table 3).
Regarding the number of phyla, each clone library from
the sediment samples of the lakes in the area of Ojos del
Salado contained members of at least 5 phyla (Table 3). It
is more than those found in similar high-altitude or permafrost-affected non-volcanic environments, for example in
the glacier melt water on Mount Everest (Liu et al. 2006),
permafrost soil of the Antarctic La Gorce mountain (Aislabie et al. 2006) and Princess Elisabeth station, Antarctica
(Peeters et al. 2011).
In the present study, the most diverse bacterial communities were detected in the samples from the lower
altitudes. Similarly diverse communities were only
reported from other active volcanic environments in the
Dry-Andes, e.g., Socompa volcano (Costello et al. 2009)
and Llullaillaco volcano (Lynch et al. 2012). Bacterial community composition of samples AS-5 and AS-6
showed the highest similarity with those inhabiting
fumaroles and soils on Socompa volcano (Costello et al.
2009), while sample AS-2 was similar to a cold fumarole
on Socompa volcano (Costello et al. 2009) and a young
volcanic sediment from lower altitude (775 m) in Japan
(Fujimura et al. 2012). This suggests that in the case of
sample AS-2 acidification due to the volcanic activity
could be one of the main environmental stress factor,
while in case of the other samples (especially AS-1, AS-5
and AS-10) the low temperature owing to the presence of
permafrost and the freeze-thaw cycles could be the main
selective factors determining the bacterial community
compositions.
AS-1, Ojos del
Salado, Dry
Andes (this
study)
AS-2, Ojos del
Salado, Dry
Andes (this
study)
AS-5, Ojos del
Salado, Dry
Andes (this
study)
AS-10, Ojos del
Salado, Dry
Andes (this
study)
AS-14, Ojos del
Salado, Dry
Andes (this
study)
AS-13, Ojos del
Salado, Dry
Andes (this
study)
Simba crater
lake, Andes
(Demergasso
et al. 2010)
Lake Namucuo,
Tibetan plateau
(Xing et al.
2009)
Socompa volcano, Andes
(Costello et al.
2009)
Socompa volcano, Andes
(Costello et al.
2009)
6500
6500
5900
5900
4350
3770
5870
4718
5824
5824
Sediment
Sediment
Sediment
Sediment
Sediment
Sediment
Sediment
Cold fumarole
Warm fumarole
Elevation
(m.a.s.l.)
Sediment
Origin of sample Sample type
+
Aquificae
+
DeinococcusThermus
+
+
+
Chloroflexi
+
Nitrospirae
+
Chalditrix
+
+
+
+
+
+
Cyanobacteria
+
Chlorobi
+
+
+
+
+
+
+
+
Alphaproteobacteria
+
+
+
+
+
+
+
+
Betaproteobacteria
Table 3 Comparison of bacterial phyla detected by clone libraries from the lake sediments of Ojos del Salado and other volcanic, high-altitude or permafrost environments
+
+
+
+
+
+
+
Gammaproteobacteria
Extremophiles (2016) 20:603–620
613
13
13
Socompa volcano, Andes
(Costello et al.
2009)
Llullaillaco
volcano, Andes
(Lynch et al.
2012)
La Gorce mountain, Antarctica
(Aislabie et al.
2006)
Princess Elisabeth Station,
Antarctica
(Peeters et al.
2011)
Volcanic sediment, Japan
(Fujimura et al.
2012)
Rongbuk glacier,
morain lake,
Mount Everest
(Liu et al.
2009)
Glacier meltwater, Mount
Everest (Liu
et al. 2009)
Salar de Aguas
Calíentes,
Andes (Demergasso et al.
2010)
Laguna Lejía,
Andes (Demergasso et al.
2010)
Laguna Vilama,
Andes (Farías
et al. 2009)
5824
6330
1800
nd
775
5140
6350
4200
4325
4650
Sediment
Sediment
Sediment
Sediment
Water
Water
Water
Water
Water
Elevation
(m.a.s.l.)
Sediment
Origin of sample Sample type
Table 3 continued
Aquificae
DeinococcusThermus
+
+
Chloroflexi
Nitrospirae
Chalditrix
+
+
Cyanobacteria
Chlorobi
+
+
+
+
+
+
Alphaproteobacteria
+
+
+
+
+
+
Betaproteobacteria
+
+
+
+
+
Gammaproteobacteria
614
Extremophiles (2016) 20:603–620
AS-1, Ojos del
Salado, Dry
Andes (this
study)
AS-2, Ojos del
Salado, Dry
Andes (this
study)
AS-5, Ojos del
Salado, Dry
Andes (this
study)
AS-10, Ojos del
Salado, Dry
Andes (this
study)
AS-14, Ojos del
Salado, Dry
Andes (this
study)
AS-13, Ojos del
Salado, Dry
Andes (this
study)
Simba crater
lake, Andes
(Demergasso
et al. 2010)
Lake Namucuo,
Tibetan plateau
(Xing et al.
2009)
Socompa volcano, Andes
(Costello et al.
2009)
Socompa volcano, Andes
(Costello et al.
2009)
Origin of sample
Table 3 continued
+
+
+
+
+
+
+
+
+
+
+
Acidobacteria
+
+
Planctomycetes
+
+
Actinobacteria
+
+
+
+
Firmicutes
+
+
Epsilone
proteobacteria
+
+
+
+
Deltaproteobacteria
+
+
+
+
+
+
+
+
+
Bacteroidetes
+
+
+
+
Verrucomicrobia
+
+
Gemmatimonadetes
+
WS5
+
+
TM7
+
OD1
Extremophiles (2016) 20:603–620
615
13
13
Deltaproteobacteria
Socompa volcano, Andes
(Costello et al.
2009)
Llullaillaco
volcano, Andes
(Lynch et al.
2012)
La Gorce mountain, Antarctica
(Aislabie et al.
2006)
Princess Elisabeth Station,
Antarctica
(Peeters et al.
2011)
Volcanic
sediment, Japan
(Fujimura et al.
2012)
Rongbuk glacier,
morain lake,
Mount Everest (Liu et al.
2009)
Glacier meltwater, Mount
Everest (Liu
et al. 2009)
Salar de Aguas
Calíentes,
Andes (Demergasso et al.
2010)
+
Laguna Lejía,
Andes (Demergasso et al.
2010)
Origin of sample
Table 3 continued
Epsilone
proteobacteria
+
+
+
+
+
+
+
+
Actinobacteria
+
+
Firmicutes
+
+
+
Planctomycetes
+
+
+
Acidobacteria
+
+
+
+
+
+
+
Bacteroidetes
+
+
+
Verrucomicrobia
+
Gemmatimonadetes
WS5
+
+
TM7
OD1
616
Extremophiles (2016) 20:603–620
Extremophiles (2016) 20:603–620
617
Conclusions
+
References
nd no data
Laguna Vilama,
Andes (Farías
et al. 2009)
Origin of sample
Table 3 continued
Deltaproteobacteria
Epsilone
proteobacteria
Firmicutes
Actinobacteria
Planctomycetes
Acidobacteria
Bacteroidetes
Verrucomicrobia
Gemmatimonadetes
WS5
TM7
OD1
The results of this study demonstrated that sediments of
high-altitude lakes located in the Ojos del Salado volcano,
Dry-Andes maintain bacterial communities with decreasing
phylogenetic diversity along with the increasing altitude.
However, the relatively high difference of bacterial diversity among the samples from the same altitude indicated
that both the extreme aridity and active volcanism influence
the bacterial community composition at the area of Ojos
del Salado. At phylum level, the phylogenetic diversity was
similar to other volcanic, high-altitude lake or permafrost
sediments from the world but at lower taxonomic levels,
representatives of several unique and potentially novel taxa
were revealed from these remote and unexplored habitats.
Abdeljabbar H, Cayol JL, Hania WB, Boudabous A, Sadfi N, Fardeau
ML (2013) Halanaerobium sehlinensesp. nov., an extremely
halophilic, fermentative, strictly anaerobic bacterium from sediments of the hypersaline lake Sehline Sebkha. Int J Syst Evol
Microbiol 63:2069–2074
Adrian R, O’Reilly CM, Zagarese H, Baines SB, Hessen DO, Keller W, Livingstone DM, Sommaruga R, Dtraile D, Van Donk E,
Weyhenmeyer GA, Winder M (2009) Lakes as sentinels of climate change. Limnol Oceanogr 54:2283
Aislabie JM, Broady PA, Saul DJ (2006) Culturable aerobic heterotrophic bacteria from high altitude, high latitude soil of La Gorce
Mountains (86 30′S, 147 W), Antarctica. Antarct Sci 18:313–321
Aislabie JM, Lau A, Dsouza M, Shepherd C, Rhodes P, Turner SJ
(2013) Bacterial composition of soils of the Lake Wellman area,
Darwin Mountains, Antarctica. Extremophiles 17:775–786
Ammann C, Jenny B, Kammer K, Messerli B (2001) Late Quaternary
glacier response to humidity changes in the arid Andes of Chile
(18–29 S). Palaeogeogr Palaeoclimatol Palaeoecol 172:313–326
Antibus DE, Leff LG, Hall BL, Baeseman JL, Blackwood CB (2012)
Cultivable bacteria from ancient algal mats from the McMurdo
Dry Valleys, Antarctica. Extremophiles 16:105–114
Avguštin JA, Bertok DŽ, Avguštin G (2013) Isolation and characterization of a novel violacein-like pigment producing psychrotrophic bacterial species Janthinobacterium svalbardensis sp.
nov. Antonie Van Leeuwenhoek 103:763–769
Azócar GF, Brenning A (2010) Hydrological and geomorphological
significance of rock glaciers in the dry Andes, Chile (27–33 S).
Permafr Periglac 21:42–53
Bajerski F, Ganzert L, Mangelsdorf K, Lipski A, Busse HJ, Padur
L, Wagner D (2013) Herbaspirillum psychrotolerans sp. nov., a
member of the family Oxalobacteraceae from a glacier forefield.
Int J Syst Evol Microbiol 63:3197–3203
Boldareva EN, Bryantseva IA, Tsapin A, Nelson K, Sorokin DY,
Tourova TP, Boichenko VA, Stadnichuk IN, Gorlenko VM
(2007) The new alkaliphilic bacteriochlorophyll a-containing
bacterium Roseinatronobacter monicus sp. nov. from the hypersaline Soda Mono Lake (California, United States). Microbiology 76:82–92
Borsodi AK, Knáb M, Krett G, Makk J, Márialigeti K, Erőss A, MádlSzőnyi J (2012) Biofilm bacterial communities inhabiting the
cave walls of the Buda Thermal Karst System, Hungary. Geomicrobiol J 29:611–627
13
618
Bowman JP, Nichols DS (2005) Novel members of the family Flavobacteriaceae from Antarctic maritime habitats including Subsaximicrobium wynnwilliamsii gen. nov., sp. nov., Subsaximicrobium saxinquilinus sp. nov., Subsaxibacter broadyi gen. nov., sp.
nov., Lacinutrix copepodicola gen. nov., sp. nov., and novel species of the genera Bizionia, Gelidibacter and Gillisia. Int J Syst
Evol Microbiol 55:1471–1486
Cabrol NA, McKay CP, Grin EA, Kiss KT, Ács E, Tóth B, Grigorszky
I, Szabó K, Fike DA, Hock AN, Demergasso C, Escudero L, Galleguillos P, Chong G, Grigsby BH,Román JZ, Tambley C (2007)
Signatures of habitats and life in Earth’s high-altitude lakes:
clues to Noachian aqueous environments on Mars. The Geology
of Mars. Evidence from earth-based analogs, pp 349–370
Cabrol NA, Grin EA, Chong G, Minkley E, Hock AN, Yu Y, Bebout L,
Fleming E, Häder DP, Demergasso C, Gibson J, Escudero L, Dorador C, Lim D, Woosley C, Morris RL, Tambley C, Gaete V, Galvez
ME, Smith E, Uskin-Peate I, Salazar C, Dawidowicz G, Majerowicz
J (2009) The high‐lakes project. J Geophys Res-Biogeo 114(G2)
Costello EK, Halloy SR, Reed SC, Sowell P, Schmidt SK (2009)
Fumarole-supported islands of biodiversity within a hyperarid,
high-elevation landscape on Socompa Volcano, Puna de Atacama, Andes. Appl Environ Microb 75:735–747
Cowan DA, Russell NJ, Mamais A, Sheppard DM (2002) Antarctic
Dry Valley mineral soils contain unexpectedly high levels of
microbial biomass. Extremophiles 6:431–436
Demergasso C, Escudero L, Casamayor EO, Chong G, Balagué V,
Pedrós-Alió C (2008) Novelty and spatio–temporal heterogeneity in the bacterial diversity of hypersaline Lake Tebenquiche
(Salar de Atacama). Extremophiles 12:491–504
Demergasso C, Dorador C, Meneses D, Blamey J, Cabrol N, Escudero L, Chong G (2010) Prokaryotic diversity pattern in high‐
altitude ecosystems of the Chilean Altiplano. J Geophys ResBiogeo 115(G2)
Dorador C, Busekow A, Vila I, Imhoff JF, Witzel KP (2008) Molecular analysis of enrichment cultures of ammonia oxidizers from
the Salar de Huasco, a high altitude saline wetland in northern
Chile. Extremophiles 12:405–414
Dorador C, Vila I, Witzel KP, Imhoff JF (2013) Bacterial and archaeal
diversity in high altitude wetlands of the Chilean Altiplano. Fund
Appl Limnol/Archiv für Hydrobiologie 182:135–159
Dunbar J, Barns SM, Ticknor LO, Kuske CR (2002) Empirical and
theoretical bacterial diversity in four Arizona soils. Appl Environ
Microb 68:3035–3045
Farías ME, Fernández‐Zenoff V, Flores R, Ordóñez O, Estévez C
(2009) Impact of solar radiation on bacterioplankton in Laguna
Vilama, a hypersaline Andean lake (4650 m). J Geophys ResBiogeo 114(G2)
Farías ME, Contreras M, Rasuk MC, Kurth D, Flores MR, Poire DG,
Novoa F, Visscher PT (2014) Characterization of bacterial diversity associated with microbial mats, gypsum evaporites and carbonate microbialites in thalassic wetlands: Tebenquiche and La
Brava, Salar de Atacama, Chile. Extremophiles 18:311–329
Finneran KT, Johnsen CV, Lovley DR (2003) Rhodoferax ferrireducens sp. nov., a psychrotolerant, facultatively anaerobic bacterium that oxidizes acetate with the reduction of Fe(III). Int J Syst
Evol Microbiol 53:669–673
Fujimura R, Sato Y, Nishizawa T, Nanba K, Oshima K, Hattori M,
Kamijo T, Ohta H (2012) Analysis of early bacterial communities on volcanic deposits on the island of Miyake (Miyake-jima),
Japan: a 6-year study at a fixed site. Microbes Environ 27:19–29
García-Moyano A, González-Toril E, Aguilera Á, Amils R (2012)
Comparative microbial ecology study of the sediments and the
water column of the Rio Tinto, an extreme acidic environment.
FEMS Microbiol Ecol 81:303–314
Goordial J, Davila A, Lacelle D, Pollard W, Marinova MM, Greer
CW, DiRuggiero J, McKay CP, Whyte LG (2016) Nearing the
13
Extremophiles (2016) 20:603–620
cold-arid limits of microbial life in permafrost of an upper dry
valley, Antarctica. ISME J. doi:10.1038/ismej.2015.239
Halloy S (1991) Islands of life at 6000 m altitude: the environment
of the highest autotrophic communities on Earth (Socompa Volcano, Andes). Arctic Alpine Res 247–262
Hammer Ř, Harper DAT, Ryan PD (2001) PAST: paleontological statistics software package for education and data analysis. Palaeontol Electron 4: 9
Hiner EJ (2009) Technical report on the Laguna Verde Salar Project,
3rd region (43-101) Copiapo. Etna Resources Inc, Chile
Hirsch P, Ludwig W, Hethke C, Sittig M, Hoffmann B, Gallikowski
CA (1998) Hymenobacter roseosalivarius gen. nov., sp. nov.
from continental Antarctic soils and sandstone: bacteria of the
Cytophaga/Flavobacterium/Bacteroides line of phylogenetic
descent. Syst Appl Microbiol 21:374–383
Hirsch P, Gallikowski CA, Siebert J, Peissl K, Kroppenstedt R, Schumann P, Stackebrandt E, Anderson R (2004) Deinococcus frigens sp. nov., Deinococcus saxicola sp. nov., and Deinococcus
marmoris sp. nov., low temperature and draught-tolerating, UVresistant bacteria from continental Antarctica. Syst Appl Microbiol 27:636–645
Houston J, Hartley AJ (2003) The central Andean west-slope rainshadow and its potential contribution to the origin of hyper-aridity in the Atacama Desert. Int J Climatol 23:1453–1464
Hultman J, Waldrop MP, Mackelprang R, David MM, McFarland J,
Blazewicz SJ, Harden J, Turetsky MR, McGuire AD, Shah MB,
VerBerkmoes NC, Lee LH, Mavrommatis K, Jansson JK (2015)
Multi-omics of permafrost, active layer and thermokarst bog soil
microbiomes. Nature 521:208–212
Kim OS, Cho YJ, Lee K, Yoon SH, Kim M, Na H, Park SC, Jeon YS, Lee
JH, Yi H, Won S, Chun J (2012) Introducing EzTaxon: a prokaryotic
16S rRNA Gene sequence database with phylotypes that represent
uncultured species. Int J Syst Evol Microbiol 62:716–721
Kim YS, Kim SJ, Anandham R, Weon HY, Kwon SW (2013)
Rhodanobacter umsongensis sp. nov., isolated from a Korean
ginseng field. J Microbiol 51:258–261
Kimura M (1980) A simple method for estimating evolutionary rates
of base substitutions through comparative studies of nucleotide
sequences. J Mol Evol 16:111–120
Krett G, Vágány V, Makk J, Jáger K, Reskóné MN, Márialigeti K,
Borsodi A (2013) Phylogenetic diversity of bacterial communities inhabiting the sediment of Lake Hévíz—a comparison of
cultivation and cloning. Acta Microbiol Imm H 60:211–235
Lane DJ (1991) 16S/23S rRNA sequencing. In: Stackebrandt E,
Goodfellow M (eds): Nucleic acid techniques in bacterial systematics. Wiley, New York, pp 115–175
Larkin MA, Blackshields G, Brown NP, Chenna R, McGettigan
PA, McWilliam H, Valentin F, Wallace IM, Wilm A, Lopez R,
Thompson JD, Gibson TJ, Higgins DG (2007) Bioinformatics
23:2947–2948. doi:10.1093/bioinformatics/btm404
Li AH, Liu HC, Xin YH, Kim SG, Zhou YG (2014) Glaciihabitans
tibetensis gen. nov., sp. nov., a psychrotolerant bacterium of the
family Microbacteriaceae, isolated from glacier ice water. Int J
Syst Evol Microbiol 64:579–587
Liu Y, Yao T, Jiao N, Kang S, Zeng Y, Huang S (2006) Microbial community structure in moraine lakes and glacial meltwaters, Mount
Everest. FEMS Microbiol Lett 265:98–105
Liu Y, Yao T, Jiao N, Kang S, Huang S, Li Q, Wang K, Liu X (2009)
Culturable bacteria in glacial meltwater at 6350 m on the East
Rongbuk Glacier, Mount Everest. Extremophiles 13:89–99
Loveland-Curtze J, Miteva VI, Brenchley JE (2009) Herminiimonas glaciei sp. nov., a novel ultramicrobacterium from
3042 m deep Greenland glacial ice. Int J Syst Evol Microbiol
59:1272–1277
Lynch RC, King AJ, Farías ME, Sowell P, Vitry C, Schmidt SK
(2012) The potential for microbial life in the highest‐elevation
Extremophiles (2016) 20:603–620
(>6000 m.a.s.l.) mineral soils of the Atacama region. J Geophys
Res Biogeo 117(G2)
Mackelprang R, Waldrop MP, DeAngelis KM, David MM, Chavarria
KL, Blazewicz SJ, Rubin EM, Jansson JK (2011) Metagenomic
analysis of a permafrost microbial community reveals a rapid
response to thaw. Nature 480:368–371
Madigan MT, Jung DO, Woese CR, Achenbach LA (2000) Rhodoferax antarcticus sp. nov., a moderately psychrophilic purple nonsulfur bacterium isolated from an Antarctic microbial mat. Arch
Microbiol 173:269–277
Männistö MK, Rawat S, Starovoytov V, Häggblom MM (2012) Granulicella arctica sp. nov., Granulicella mallensis sp. nov., Granulicella tundricola sp. nov. and Granulicella sapmiensis sp. nov.,
novel acidobacteria from tundra soil. Int J Syst Evol Microbiol
62:2097–2106
Männistö MK, Tiirola M, McConnell J, Häggblom MM (2010) Mucilaginibacter frigoritolerans sp. nov., Mucilaginibacter lappiensis
sp. nov. and Mucilaginibacter mallensis sp. nov., isolated from
soil and lichen samples. Int J Syst Evol Microbiol 60:2849–2856
Margesin R, Spröer C, Zhang DC, Busse HJ (2012) Polaromonas
glacialis sp. nov. and Polaromonas cryoconiti sp. nov., isolated from alpine glacier cryoconite. Int J Syst Evol Microbiol
62:2662–2668
Massol-Deya AA, Odelson DA, Hickey RF, Tiedje JM (1995) Bacterial community fingerprinting of amplified 16S and 16–23S
ribosomal DNA gene sequences and restriction endonuclease
analysis (ARDRA). In: Molecular microbial ecology manual.
Springer, The Netherlands, pp 289–296
Messing J (1983) New M13 vectors for cloning. Method Enzymol
101:20–78
Muyzer G, De Waal EC, Uitterlinden AG (1993) Profiling of complex
microbial populations by denaturing gradient gel electrophoresis
analysis of polymerase chain reaction-amplified genes coding for
16S rRNA. Appl Environ Microb 59:695–700
Nagy B, Mari L, Kovács J, Nemerkényi Zs, Heiling Zs, (2014/a)
Környezetváltozás a Száraz-Andokban: az Ojos del Salado
monitoring vizsgálata (In Hungarian).In: Cserny T, KovácsPálffy P, Krivánné Horváth Á (szerk.) HUNGEO 2014 Magyar Földtudományi szakemberek XII. találkozója: Magyar
felfedezők és kutatók a természeti erőforrások hasznosításáért:
cikkgyűjtemény. Budapest: Magyarhoni Földtani Társulat, pp
53–62 (ISBN:978-963-8221-53-7)
Nagy B, Mari L, Kovács J, Nemerkényi Zs, Heiling Zs (2014b) Az
Ojos del Salado monitoring vizsgálata: jég- és vízjelenlét a Föld
legszárazabb magashegységében (In Hungarian).In: Sansumné
Molnár J, Siskáné Szilasi B, Dobos E (szerk) VII. Magyar Földrajzi Konferencia kiadványa, Miskolc, pp 449–459. (ISBN:
978-963-358-063-9)
Nakagawa Y, Sakane T, Suzuki M, Hatano K (2002) Phylogenetic
structure of the genera Flexibacter, Flexithrix, and Microscilla
deduced from 16S rRNA sequence analysis. J Gen Appl Microbiol 48:155–165
Nakai R, Abe T, Baba T, Imura S, Kagoshima H, Kanda H, Kanekiyo
A, Kohara Y, Koi A, Nakamura K, Narita T, Niki H, Yanagihara
K, Naganuma T (2012) Microflorae of aquatic moss pillars in
a freshwater lake, East Antarctica, based on fatty acid and 16S
rRNA gene analyses. Polar Biol 35:425–433
Nübel U, Engelen B, Felske A, Snaidr J, Wieshuber A, Amann
RI, Ludwig W, Backhaus H (1996) Sequence heterogeneities of genes encoding 16S rRNAs in Paenibacillus polymyxa
detected by temperature gradient gel electrophoresis. J Bacteriol
178:5636–5643
Ordoñez OF, Flores MR, Dib JR, Paz A, Farías ME (2009) Extremophile culture collection from Andean lakes: extreme pristine
environments that host a wide diversity of microorganisms with
tolerance to UV radiation. Microb Ecol 58:461–473
619
Peeters K, Ertz D, Willems A (2011) Culturable bacterial diversity at the Princess Elisabeth Station (Utsteinen, Sør Rondane
Mountains, East Antarctica) harbours many new taxa. Syst Appl
Microbiol 34:360–367
Ruckmani A, Kaur I, Schumann P, Klenk HP, Mayilraj S (2011)
Calidifontibacter indicus gen. nov., sp. nov., a member of the
family Dermacoccaceae isolated from a hot spring, and emended
description of the family Dermacoccaceae. Int J Syst Evol
Microbiol 61:2419–2424
Sattin SR, Cleveland CC, Hood E, Reed SC, King AJ, Schmidt SK,
Robeson MS, Ascarrunz N, Nemergut DR (2009) Functional
shifts in unvegetated, perhumid, recently-deglaciated soils do not
correlate with shifts in soil bacterial community composition. J
Microbiol 47:673–681
Schumann P, Zhang DC, Redzic M, Margesin R (2012) Alpinimonas
psychrophila gen. nov., sp. nov., an actinobacterium of the family Microbacteriaceae isolated from alpine glacier cryoconite. Int
J Syst Evol Microbiol 62:2724–2730
Scott S, Dorador C, Oyanedel JP, Tobar I, Hengst M, Maya G, Harrod
C, Vila I (2015) Microbial diversity and trophic components of
two high altitude wetlands of the Chilean Altiplano/Diversidad
microbiana y componentes trófi cos de dos humedales de altura
del altiplano chileno. Gayana 79:45
Smalla K, Oros-Sichler M, Milling A, Heuer H, Baumgarte S, Becker
R, Neuber G, Kropf S, Ulrich A, Tebbe CC (2007) Bacterial
diversity of soils assessed by DGGE, T-RFLP and SSCP fingerprints of PCR-amplified 16S rRNA gene fragments: do the
different methods provide similar results? J Microbiol Methods
69:470–479
Stackebrandt E, Goebel BM (1994) Taxonomic note: a place for
DNA–DNA reassociation and 16S rRNA sequence analysis in
the present species definition in bacteriology. Int J Syst Evol
Microbiol 44:846–849
Subcommittee Permafrost (1988) Glossary of permafrost and related
ground-ice terms. Associate Committee on Geotechnical
Research, National Research Council of Canada, Ottawa, p 156
Tamura K, Stecher G, Peterson D, Filipski A, Kumar S (2013)
MEGA6: molecular evolutionary genetics analysis version 6.0.
Mol Biol Evol 30:2725–2729
Taş N, Prestat E, McFarland JW, Wickland KP, Knight R, Berhe AA,
Jorgeson T, Waldrop MP, Jansson JK (2014) Impact of fire on
active layer and permafrost microbial communities and metagenomes in an upland Alaskan boreal forest. ISME J 8:1904–1919
Tindall BJ, Rosselló-Mora R, Busse HJ, Ludwig W, Kämpfer P
(2010) Notes on the characterization of prokaryote strains for
taxonomic purposes. Int J Syst Evol Microbiol 60:249–266
Van Trappen S, Vandecandelaere I, Mergaert J, Swings J (2005)
Flavobacterium fryxellicola sp. nov. and Flavobacterium psychrolimnae sp. nov., novel psychrophilic bacteria isolated from
microbial mats in Antarctic lakes. Int J Syst Evol Microbiol
55:769–772
Vandamme P, Pot B, Gillis M, De Vos P, Kersters K, Swings J (1996)
Polyphasic taxonomy, a consensus approach to bacterial systematics. Microbiol Rev 60:407–438
Vuille M, Ammann C (1997) Regional snowfall patterns in the high,
arid Andes. Clim Chang 36:413–423
Watson SW, Bock E, Valois FW, Waterbury JB, Schlosser U (1986)
Nitrospira marina gen. nov. sp. nov.: a chemolithotrophic nitriteoxidizing bacterium. Arch Microbiol 144:1–7
Wu QL, Zwart G, Schauer M, Kamst-van Agterveld MP, Hahn
MW (2006) Bacterioplankton community composition along
a salinity gradient of sixteen high-mountain lakes located
on the Tibetan Plateau, China. Appl Environ Microbiol
72:5478–5485
Xing P, Hahn MW, Wu QL (2009) Low taxon richness of bacterioplankton in high-altitude lakes of the eastern Tibetan Plateau,
13
620
with a predominance of Bacteroidetes and Synechococcus spp.
Appl Environ Microbiol 75:7017–7025
Zhang DC, Wang HX, Liu HC, Dong XZ, Zhou PJ (2006) Flavobacterium glaciei sp. nov., a novel psychrophilic bacterium
isolated from the China no. 1 glacier. Int J Syst Evol Microbiol
56:2921–2925
Zhang G, Ma X, Niu F, Dong M, Feng H, An L, Cheng G (2007)
Diversity and distribution of alkaliphilic psychrotolerant bacteria
13
Extremophiles (2016) 20:603–620
in the Qinghai-Tibet Plateau permafrost region. Extremophiles
11:415–424
Zhang DC, Schumann P, Redzic M, Zhou YG, Liu HC, Schinner F,
Margesin R (2012) Nocardioides alpinus sp. nov., a psychrophilic actinomycete isolated from alpine glacier cryoconite. Int J
Syst Evol Microbiol 62:445–450
Extremophiles is a copyright of Springer, 2016. All Rights Reserved.
Descargar