You are on page 1of 16

bs_bs_banner

Environmental Microbiology (2015) 17(10), 3882–3897 doi:10.1111/1462-2920.12872

Melting glacier impacts community structure of


Bacteria, Archaea and Fungi in a Chilean
Patagonia fjord

Marcelo H. Gutiérrez,1,2* Pierre E. Galand,3 Introduction


Carlos Moffat1,2,4 and Silvio Pantoja1,2
1
Most glaciers in Patagonian Ice Fields have retreated
Department of Oceanography,
2
over the past half-century (Rignot et al., 2003; Willis et al.,
COPAS Sur-Austral Program, Universidad de
2012), with annual ice mass losses of 24.4 ± 1.4 Gt, which
Concepción, Concepción, Chile.
3
are equivalent to a sea level rise of 0.067 ± 0.004 mm
Sorbonne Universités, UPMC Univ Paris 06, CNRS,
year−1 (Willis et al., 2012). Jorge Montt glacier (Fig. 1), the
Laboratoire d’Ecogéochimie des Environnements
northern-most glacier in the Southern Patagonian Ice
Benthiques (LECOB), Observatoire Océanologique,
Field, has experienced a rapid retreat of c. 19.5 km
F-66650, Banyuls sur Mer, France.
4
between 1898 and 2011, and an accelerated rate of ter-
Institute of Marine Sciences, University of California
minal ice loss in recent years (Rivera et al., 2012). Calving
Santa Cruz, Santa Cruz, CA, USA.
of the glacier front as well as submarine melting and
run-off result in an input of freshwater that impacts the
Summary hydrographic structure and water circulation in the
proglacial fjord (Moffat, 2014).
Jorge Montt glacier, located in the Patagonian Ice
The hydrographic structure of Patagonian fjords and
Fields, has undergone an unprecedented retreat
channels in southern Chile is characterized by two or
during the past century. To study the impact of the
more layers originating from high freshwater discharge
meltwater discharge on the microbial community of
from terrestrial drainage, rivers, precipitation and ice
the downstream fjord, we targeted Bacteria, Archaea
melting, and by the deep input of Subantarctic waters
and Fungi communities during austral autumn and
from the adjacent Pacific Ocean (Silva et al., 1998; Dávila
winter. Our results showed a singular microbial com-
et al., 2002). The freshwater flux supplies allochthonous
munity present in cold and low salinity surface waters
organic and inorganic matter to these fjords (Sepúlveda
during autumn, when a thicker meltwater layer was
et al., 2011; Silva et al., 2011; Vargas et al., 2011;
observed. Meltwater bacterial sequences were related
González et al., 2013), which may influence the plank-
to Cyanobacteria, Proteobacteria, Actinobacteria and
tonic trophic web (Vargas et al., 2011), and autotrophic
Bacteriodetes previously identified in freshwater and
and heterotrophic microbial activity (Montero et al., 2011).
cold ecosystems, suggesting the occurrence of
In particular, meltwater represents a sizable source of
microorganisms adapted to live in the extreme con-
freshwater (Rignot et al., 2003; Willis et al., 2012) charac-
ditions of meltwater. For Fungi, representative
terized by high load of suspended sediment (Boldt et al.,
sequences related to terrestrial and airborne fungal
2013), high concentration of silicic acid and low nitrate
taxa indicated transport of allochthonous Fungi by
and phosphate content (González et al., 2011; 2013) that
the meltwater discharge. In contrast, bottom fjord
can impact primary productivity and the structure of
waters from autumn and winter showed representa-
pelagic community in the fjord system adjacent to the ice
tive Operational Taxonomic Units (OTUs) related to
fields (González et al., 2013).
sequences of marine microorganisms, which is con-
The microbial community in cold aquatic environments
sistent with current models of fjord circulation. We
includes representative members of the domains Bacte-
conclude that meltwater can significantly modify the
ria, Archaea and Eukarya (Margesin and Miteva, 2011).
structure of microbial communities and support the
Among Bacteria, Alphaproteobacteria, Betaproteo-
development of a major fraction of microorganisms in
bacteria, Cyanobacteria, Bacteriodetes and Actinoba-
surface waters of Patagonian fjords.
cteria frequently predominate in cryoconites holes
(Edwards et al., 2011; Cameron et al., 2012), snow
Received 20 October, 2014; revised 2 April, 2015; accepted 2 April,
2015. *For correspondence. E-mail magutier@udec.cl; Tel. (+56) 41 and glacial meltwater (Møller et al., 2013), glacier fed
266 1287; Fax (+56) 41 222 5400. streams (Wilhelm et al., 2013), high mountain glaciers
© 2015 Society for Applied Microbiology and John Wiley & Sons Ltd
Microbial community in meltwaters of Patagonian fjords 3883

Fig. 1. Study area and sampling sites during autumn (red circles) and winter (yellow triangles) in the fjord adjacent to Jorge Montt glacier with
a colour scale of bathymetry (A). Salinity sections (coloured scale) and temperature isolines (black trace) for the April (B) and August (C) 2012
field campaigns.

(Liu et al., 2011) and glacial fjords (Zeng et al., 2009). Patagonian glaciers are retreating (Rignot et al., 2003;
Among Archaea, Thaumarchaeota has been detected in Rivera et al., 2007; Willis et al., 2012) and since meltwater
Antarctic cryoconite holes (Cameron et al., 2012) and has the potential to change the physicochemical environ-
sea ice (Cowie et al., 2011). Although the composition of ment of the downstream ecosystems, we hypothesize that
the fungal community remains less known in cold envi- melting impacts the microbial community composition in
ronments, yeast, melanized Fungi and filamentous Patagonian proglacial fjords by favouring the develop-
Ascomycota have been found in supraglacial systems ment of meltwater-adapted microorganisms. In order to
(Uetake et al., 2012; Edwards et al., 2013) and in Arctic test our hypothesis, we analysed the spatial (vertical and
ice (Gunde-Cimerman et al., 2003). In fjords and high horizontal) and seasonal variability of Bacteria, Archaea
latitude waters influenced by glacial meltwater, members and Fungi communities in the water column and sedi-
of classes Gammaproteobacteria and Alphaproteo- ments of the fjord adjacent to Jorge Montt glacier.
bacteria and the phylum Cytophaga–Flavobacterium–
Bacteroides frequently dominate bacterial communities Results
(Zeng et al., 2009; 2013; Piquet et al., 2010; 2011).
Environmental variability
Moreover, changes in hydrographic properties as a result
of meltwater input can induce temporal and spatial During autumn (April), cold (<4°C) and fresh (<20 psu)
changes in the structure of the microbial community surface water occupied the top 10 m of the water column
(Piquet et al., 2010; Zeng et al., 2013). Differences in near the glacier terminus (Fig. 1B). The layer thinned and
bacterial composition have been observed associated to its temperature and salinity increased from the terminus
varying meltwater inflow (Piquet et al., 2010) and between towards the mouth of the fjord connecting to Baker
fjord waters and the adjacent ocean (Zeng et al., 2009). Channel (c. 20 km, Fig. 1B). Below this layer, temperature
Among photosynthetic microorganisms, meltwater inflow increased gradually from 6°C to c. 10°C in bottom waters
seems to favour the occurrence of small size nano- and of the fjord basin and c. 9°C in Baker Channel (Fig. 1B).
picophytoplankton (Piquet et al., 2014). Salinity also showed a gradual increase from c. 24 to

© 2015 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 17, 3882–3897
3884 M. H. Gutiérrez, P. E. Galand, C. Moffat and S. Pantoja
Table 1. Sample ID and chemical and biological characteristics of the waters sampled during April and August 2012 in the fjord adjacent to Jorge
Montt glacier.

Fungal
Sample Depth Nitrate Phosphate Silicic acid Chlorophyll-α Prokaryote filaments
Date ID Station Location (m) (μM) (μM) (μM) (mg m−3) (×108 cell l−1) (ng C l−1)

April 2012 A2-3m A2 48°15.8’S 3 n.d. 0.01 ± 0.01 16.1 ± 0.4 0.14 3.2 ± 0.2 224.7 ± 219.0
73°26.7’W
A3-3m A3 48°14.8 S 3 5.3 ± 1.0 n.d. 18.2 ± 0.7 0.29 1.6 ± 0.0 38.8 ± 40.7
73°26.7 W
A4-3m A4 48°13.9′ S 3 n.d. n.d. 24.5 ± 0.4 0.32 4.0 ± 0.0 28.9 ± 33.6
73°28.2′ W
A5-3m A5 48°06,6′ S 3 2.5 ± 0.6 0.2 ± 0.01 9.5 ± 1.8 0.10 3.0 ± 0.2 10.3 ± 13.8
73°29,7′ W
A1-210m A1 48°15.9’S 210 7.6 ± 0.4 0.8 ± 0.01 3.3 ± 0.1 n.d. 1.1 ± 0.2 n.d.
773°26.9′ W
A2-290m A2 48°15.8’S 290 8.0 ± 0.1 1.0 ± 0.02 6.2 ± 0.2 n.d. 1.6 ± 0.0 n.d.
73°26.7’W
August 2012 W1-0m W1 48°17,8′ S 0 0.8 ± 0.0 0.08 ± 0.00 20.6 ± 0.0 0.02 2.9 ± 0.0 95.2 ± 4.1
73°27,3′ W
W2-0m W2 48°14,2′ S 0 1.0 ± 0.0 0.09 ± 0.00 25.8 ± 0.0 0.6 2.6 ± 0.1 71.8 ± 17.0
73°28,2′ W
W5-3m W5 48°08,8′ S 3 1.4 ± 0.0 0.12 ± 0.00 29.2 ± 0.0 0.34 2.4 ± 0.0 253.8 ± 228.3
73°26,1′ W
W1-300m W1 48°17,8′ S 300 17.7 ± 0.0 0.94 ± 0.00 6.21 ± 0.0 n.d. 1.2 ± 0.0 194.1 ± 115.3
73°27,3′ W
W2-50m W2 48°14,2′ S 50 11.0 ± 0.0 1.05 ± 0.00 4.6 ± 0.0 n.d. 1.2 ± 0.0 6.1 ± 4.1
73°28,2′ W
W5-100m W5 48°08,8′ S 100 24.7 ± 0.0 1.40 ± 0.00 12.5 ± 0.0 n.d. 1.0 ± 0.0 71.8 ± 17.0
73°26,1′ W
W3-sed W3 48°11,6′ S Surf. sed.
73°30,2′ W
W4-sed W4 48°11,5′ S Surf. sed.
73°29,3′ W

Concentrations of nutrients and abundance of prokaryote and biomass of fungal filaments are average ± standard deviation of duplicate samples.
n.d., not detected.

30 psu in waters in the fjord basin and maximum values to 10°C and 29 to over 32 psu in the inner fjord and in Baker
(c. 34 psu) in the deeper waters of Baker channel Channel (Fig. 1C). In this period, the salinity below 15 m in
(Fig. 1B). The hydrographic structure within the fjord is the fjord basin and in Baker Channel was higher than
strongly influenced by the presence of a sill located at the 30 psu. A subsurface temperature maximum was found
mouth of the fjord that blocks the inflow of water below deeper in Baker Channel (at around 45 m depth) and had
60 m deep from Baker Channel (Fig. 1A). On the distal a maximum value roughly one degree lower (c. 10.5°C)
side of the sill, water column was characterized by a than in autumn (Fig. 1B and C).
subsurface temperature maximum of c. 11.5°C centered Surface waters had lower nitrate and phosphate con-
at 30 m, while from the sill and inward, water was colder centrations than bottom waters during autumn and winter
and fresher, a result of mixing with the fresh and cold (Table 1). In contrast, higher concentrations of silicic acid
meltwater. A tongue of warm and salty water was were observed in surface relative to bottom waters
observed intruding at depth from the sill, which is located (Table 1). Chlorophyll-α in the surface layer showed an
around 35 m along this section and reaches a maximum average concentration of 0.17 ± 0.09 mg m−3 in autumn
depth of 45 m (Fig. 1B). and 0.30 ± 0.22 mg m−3 in winter (Table 1), and no major
During winter (August), the fresh surface layer was differences were observed from the head to the mouth of
thinner than in autumn, with the halocline located at c. 5 m the fjord. In addition, prokaryote abundance and fungal
deep (Fig. 1C). Temperature and salinity in the surface biomass were higher in surface than bottom waters during
layer was not homogenous, with colder (<3°C) and fresher both seasons (Table 1). There were significant differences
(<15 psu) water located close to the glacier front and at the in silicic acid concentration at different depths and during
mouth of the fjord, the latter suggesting an intrusion of both seasons as well as in nitrate and phosphate concen-
freshwater from Baker Channel (Fig. 1C). Both tempera- trations and prokaryote abundance between surface and
ture and salinity increased gradually with depth from c. 5°C bottom waters (Table 2).

© 2015 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 17, 3882–3897
Microbial community in meltwaters of Patagonian fjords 3885
Table 2. Two-way ANOVA to test statistical differences on the rank- Fungi (Fig. S1) indicated that all samples had reached the
transformed data of environmental variables and bacterial and fungal
curvilinear phase of sampling effort, and that for Archaea,
richness between seasons and depths.
most of the curves had started to plateau.
Variable Factor df F P-value OTU richness estimated by Chao1 after resampling
ranged from 57 to 367 for Bacteria, 1 to 97 for Archaea
Nitrogen Period 1 0.00 1.000
Depth 1 58.3 0.000 and 5 to 146 for Fungi (Table 3). In 12 out of 14 samples,
Period×Depth 1 18.0 0.005 the observed number of bacterial and fungal OTUs rep-
Phosphate Period 1 0.6 0.467 resented over 80% of the OTU richness estimated by
Depth 1 20.2 0.004
Period×Depth 1 0.6 0.468 Chao1 (Table 3). For Archaea, the abundance detected in
Silicic acid Period 1 5.45 0.048 samples with a relative high number of OTUs (>20) rep-
Depth 1 30.21 0.000 resented over 70% of the richness estimated by Chao1.
Period×Depth 1 0.28 0.612
Prokaryote Period 1 1.12 0.321 Bacterial diversity (Chao1, Table 3) was higher in the
Depth 1 19.18 0.002 bottom (saline) than the surface waters (low salinity) of the
Period×Depth 1 0.04 0.837 fjord basin in autumn and winter (Stations A1-A4 and
Chlorophyll – – – –
Fungal filaments – – – – W1-W2; Table 3). In contrast, fungal richness in the fjord
Chao1 Bacteria Period 1 3.31 0.106 basin was higher in the surface than the bottom layer in
Depth 1 0.81 0.395 both periods (Table 3). Archaeal diversity was highest in
Period×Depth 1 0.18 0.686
Chao1 Archaea Period 1 1.09 0.328 sediments collected during winter close to the fjord mouth
Depth 1 1.46 0.262 and in surface waters from Baker channel (stations W3
Period×Depth 1 0.92 0.365 and W4, Table 3). There were no significant differences
Chao1 Fungi Period 1 0.43 0.530
Depth 1 0.61 0.456 (at 95% confidence level) in bacterial, fungal and archaeal
Period×Depth 1 4.58 0.065 diversity between seasons and depths (Table 2).
Bacterial diversity in surface waters in both autumn and
–, insufficient data for analysis. The values highlighted in bold are
statistically significant at 95% confidence level.
winter was higher in the Baker Channel (sites A5 and W5)
than in the fjord basin (sites A2-A4 and W1 and W2;
Table 3). Fungal communities were more diverse in
Bacteria, Archaea and Fungi community diversity
surface waters in the fjord basin (sites A2-A4) than in
Microbial diversity was analysed by comparing the distri- Baker Channel (site A5) in autumn. In winter, Baker
bution of 34 263 and 21 692 16S rRNA gene sequences of Channel waters had the highest fungal diversity (Table 3).
Bacteria and Archaea respectively, and 38 891 ITS rDNA OTU richness for Bacteria and Fungi showed higher
sequences of Fungi (Table 3). A total of 1443 Operational values during austral autumn (April) compared with winter
Taxonomic Units (OTUs) of Bacteria, 164 of Archaea and (August) in surface waters (Table 3). In bottom waters in
298 of Fungi were identified. Differences in OTU abun- the fjord basin (sites A1, A2, W1 and W2), diversity was
dances among depths, sampling sites and seasons were higher in April for bacteria and in August for Fungi
observed (Table 3). Rarefaction curves for Bacteria and (Table 3). Additionally, in August the highest Bacteria,

Table 3. Number of sequences, OTU abundance and richness after resampling (Chao1) estimated for microbial assemblages from waters and
sediments of the fjord adjacent to Jorge Montt glacier.

Bacteria Archaea Fungi

Sample ID Reads OTUs Chao1 Reads OTUs Chao1 Reads OTUs Chao1

A2-3m 1999 184 167 1402 136 146


A3-3m 2790 199 186 60 3 1894 73 70
A4-3m 5836 285 104 2767 8 1 1085 97 99
A5-3m 3038 342 252 1369 22 17
A1-210m 1865 324 300 18 7 1827 20 18
A2-290m 1106 251 188 378 21 23 6361 26 18
W1-0m 627 47 57 1277 8 4 3549 33 25
W2-0m 347 54 68 1998 9 10 6709 56 23
W5-3m 1872 310 211 2340 82 42 1951 36 41
W1-300m 2835 138 92 8 3 760 42 53
W2-50m 1825 124 90 1855 33 47 4765 47 35
W5-100m 5559 376 201 7031 60 25 1465 83 82
W3-sed 4564 589 367 2264 122 97 1645 54 53
W4-sed 1696 3 2 4109 10 5

© 2015 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 17, 3882–3897
3886 M. H. Gutiérrez, P. E. Galand, C. Moffat and S. Pantoja

Archaea and Fungi diversity was observed in surface Cluster B1 was mainly characterized by OTUs matching
sediment collected near the fjord mouth (site W3, c. 38 m clones of uncultured Synechococcus, Comamonadaceae
depth, Table 3). Betaproteobacteria, Alphaproteobacteria and Actino-
bacteria of the group Actinomycetales (Fig. 2B, Table 4).
In this group (surface waters in autumn), Cyanobacteria
Composition of bacterial communities dominated in waters in the fjord basin (more than 40% of
Bacterial communities were separated according to their the total Bacteria sequences), followed by Proteobacteria,
composition at the OTU level into three major clusters with Actinobacteria and Bacteroidetes (Fig. 2C). In contrast,
a similarity level (based on a Bray–Curtis distance matrix) in surface waters of Baker Channel (sample A5-3m)
higher than 70%. Cluster B1 grouped the surface water sequences of Proteobacteria were the most abundant
community from autumn, B2 included the bottom water (50%) followed by Cyanobacteria (c. 25%; Fig. 2C).
community from winter and B3 the bottom water commu- Among Proteobacteria (Fig. 2D), sequences of Alphapro-
nity from autumn (Fig. 2A). There were two exceptions to teobacteria predominated in surface waters in the fjord
this grouping: sites W2-0m in cluster B2 and W5-3m in basin (36–44% of total Proteobacteria sequences) and
cluster B3. The clustering was consistent with the pattern Gammaproteobacteria in surface waters from Baker
observed in the ordination plots derived from principal Channel (site A5, 41% of Proteobacteria sequences;
coordinate analysis (PCoA; Fig. S2). Fig. 2D).

Fig. 2. Dendrogram based on Bray–Curtis similarity between bacterial assemblages in surface and bottom waters from autumn and winter in
the fjord adjacent to Jorge Montt glacier (A). Heatmap of the main bacterial OTUs (B), relative abundance (percentage of total sequences) of
the main bacterial groups (C) and detail for Proteobacteria (D). B1, B2 and B3 represent the bacterial communities identified with more than
70% of similarity.

© 2015 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 17, 3882–3897
Microbial community in meltwaters of Patagonian fjords 3887
Table 4. Best taxonomic match for the main OTUs of Bacteria, Fungi and Archaea communities in the water column and sediment of the fjord
adjacent to Jorge Montt glacier.

Representative Accession
OTU ID cluster/sample Best taxonomic match number % identity Environment Reference

BacJM1 B1 Synechococcus AY151250 99 Subalpine and high altitude lakes Crosbie et al., 2003a, b; Ernst
EU703201 et al., 2003; Xing et al., 2009
EU703437
EU703201
BacJM2 B1, B2 Comamonadaceae DQ234234 100 Estuarine waters Liao et al., 2007
Unclassified Bacteria JF429247 River waters Liu et al., 2012
BacJM3 B1, B2 Alphaproteobacteria DQ063213 98 Coastal environment influenced by Riemann et al., 2008
freshwater discharge
BacJM4 B2 Unclassified Bacteria HQ730061 99 Bacteria associated to Antarctic Murray et al., 2011
icebergs
BacJM5 B1 Actinomycetales HM856562 97 Yellowstone lake Clingenpeel et al., 2011
BacJM8 B2 Unclassified Bacteria EU287086 100 Surface sediments from the Pacific Li et al., 2009
FR685473 Arctic Ocean Newbold et al., 2012
Fjord waters
BacJM9 B3 Unclassified Bacteria EU491101 97–98 Ocean crust and deep sea mud Santelli et al., 2008;
HQ588368 96 volcanoes Pachiadaki et al., 2011
FJ640811 95 Hydrothermal vents Wang et al., 2009
JF344181 Subtidal sediments Acosta-Gonzalez et al., 2013
BacJM10 B2 Alphaproteobacteria of the RSU63934 99 Waters of the Baltic Sea Pinhassi et al., 1997
genus Rhodobacter
BacJM11 B2 Unclassified Bacteria DQ015848 97 Antarctic lakes Glatz et al., 2006; Tang et al.,
Bacteriodetes JX948651 95 Subantarctic waters 2013
(Flavobacteriales) AY794159 Prabagaran et al., 2007
BacJM13 B3 Unclassified Bacteria EU035849 100 Deep-water coral reefs Jensen et al., 2008
HQ673544 99 Subarctic waters Allers et al., 2013
HQ729998 99 Associated to Antarctic icebergs Murray et al., 2011
BacJM19 B2 Unclassified bacteria EU919855 99 Glacial Arctic fjord Zeng et al., 2009
BacJM37 B3, W3-sed Alphaproteobacteria KC492638 100 Sponge-associated bacterial Croué et al., 2013
Unclassified bacteria KC197685 100 community Sjöstedt et al., 2014
Marine environments
BacJM51 W1-0m Gammaproteobacteria (Vibrio) KF942293 100 Upwelling region, Monterey Bay Mansergh and Zehr, 2014
BacJM70 W1-0m, B2 Unclassified bacteria FJ628262 100 Anoxic tidal fjord Schmidtova et al., 2009
EF379591 98 Estuarine environment Zhang et al., 2007
BacJM182 W1-0m Gammaproteobacteria (Vibrio) KF941811 100 Upwelling region, Monterey Bay Mansergh and Zehr, 2014
FunJM1 F1, F2, F3 Basidiomycota GU062301 100 Indoor dust and associated to Arhipova et al., 2011, Wirsel
Unclassified Fungi clone AJ279465 100 terrestrial plants et al., 2001
FJ820495 Air particulate matter Fröhlich-Nowoisky et al., 2009
FunJM2 F4, F1 Cladosporium (Ascomycota) KF225878 100 Soil Hu et al., 2013
KF367474 Freshwater Oliveira et al., 2013
HE608784 Coastal environment Scopel et al., 2013
FunJM3 F2, F3 Basidiomycota, Agaricales KC176337 95–97 Soils Thorn et al., 1996;
KC669316 Ramírez-Cruz et al., 2013
FunJM4 W4-sed Cryptococcus (Basidiomycota) DQ643977 100 In association with mycorrhizal Alonso et al., 2008
DQ317387 fungi Arenz et al., 2006
Antarctic soils
FunJM5 W1-0m Unclassified fungi JF945133 <90 Forest environment Cordier et al., 2012
FunJM6 F1, F2, F3 Cryptococcus (Basidiomycota) FR717837 99 Arbuscular mycorrhizal roots and Yurkov et al., 2012
AJ581048 spores Renker et al., 2004
FunJM7 F3, F4 Unclassified fungi FJ235861 99 Soils Hong et al., 2010
FunJM8 W4-sed Unclassified fungi JF497138 99 Deep-Sea sediments Singh et al., 2012
FunJM9 F1 Unclassified Fungi KF296942 <90 Arctic soils Timling et al., 2014
FunJM10 A5-3m Chytridiomycete HQ191313 95 Lake ecosystem Monchy et al., 2011
FunJM14 F3 Unclassified fungi AB615542 99 Subsurface marine sediments Y. Nagano, T. Nagahama, M.
Konishi, T. Kubota, F. Abe,
K. Takahashi, Y. Hatada,
unpublished
FunJM18 F1, F2, F4 Unclassified Fungi KC965831 100 Arctic soils Timling et al., 2014
Cladosporium (Ascomycota) HM148149 Cosmopolitan fungi Bensch et al., 2010
KF156274 Root associated fungi Stenström et al., 2013
FunJM21 F3 Penicillium (Ascomycota) GU062216 100 Associated with plants Arhipova et al., 2011
FunJM23 F1, A5-3m Hypocreales (Ascomycota) 99 Root associated fungi in saline Macia-Vicente et al., 2012
gradient
FunJM57 F1 Unclassified Fungi JX974768 <90 Estuarine sediments B. Xing, X. Zhang, J. Gong,
unpublished
ArcJM1 A4-3m, W2-50m, W5-100m Unclassified Archaea HE796570 99 Freshwater environments including Auguet and Casamayor, 2012;
KC604531 high mountain lakes Flynn et al., 2013; Kato
AB722183
et al., 2013
ArcJM2 W2-50m, W5-100m, W3-sed, Thaumarchaeota EF069371 100 Antarctic sediments Gillan and Danis, 2007
W2-0m, W5-3m, W4-sed
ArcJM4 W3-sed, W5-3m, W1-0m Thaumarchaeota, Marine JX015356 100 Meromitic saline lake La Cono et al., 2013
group I GQ387725 Mediterranean surface waters Galand et al., 2010

© 2015 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 17, 3882–3897
3888 M. H. Gutiérrez, P. E. Galand, C. Moffat and S. Pantoja

Waters from August (cluster B2, Fig. 2) shared main Ascomycota sequences accounted for between 36% and
bacterial OTUs with the microbial community from the 66% and Basidiomycota from 26% to 50% of the fungal
surface meltwater in April (cluster B1; Fig. 2, Table 4). In sequences (Fig. 3C). In the same cluster, sequences of
addition, OTUs matching clones of unclassified Bacteria, unclassified fungal OTUs accounted for 7–18% of the
Alphaproteobacteria of the genus Rhodobacter and total sequences (Fig. 3C).
Bacteriodetes, were also representative of this community In F2 (waters from August), the community was
(Table 4, Fig. 2). Cluster B2 (mostly bottom waters from characterized by OTUs related to clones of Basidio-
August) was also characterized by a predominance of mycota of the order Agaricales, uncharacterized Fungi
Proteobacteria reaching c. 50% of the total bacterial and Cladosporium spp. (Fig. 3; Table 4). Basidiomycota
sequences, followed by Cyanobacteria and Bacteriodetes sequences represented more than 87% and 50% of the
(Fig. 2C). Among Proteobacteria, sequences of Alphapro- sequences in surface and bottom waters respectively
teobacteria predominated (40–54% of total Proteoba- (Fig. 3C).
cteria sequences) followed by Betaproteobacteria and Cluster F3 shared main OTUs with F2 and included
Gammaproteobacteria (Fig. 2D). The bacterial community OTUs matching Ascomycota of the genus Penicillium and
in surface waters from site W2 was also included in cluster uncharacterized Fungi (Fig. 3B; Table 4). This community
B2. (water from 100 m depth in Baker Channel and sediment
Cluster B3 was mainly composed of two OTUs match- close to the fjord mouth in winter) was characterized by a
ing clones of an unclassified Bacteria and an Alpha- predominance of Ascomycota Fungi sequences (46–
proteobacteria from the marine environment (Fig. 2B, 53%; Fig. 3C). In addition, sequences of Chytridiomycota
Table 4). The bacterial community grouped in B3 (bottom represented c. 8% in waters from Baker Channel (100 m)
waters from April) was dominated by sequences of and unclassified fungal sequences accounted for 24% in
Proteobacteria representing more than 70% of bacterial surface sediments (Fig. 3C).
sequences (Fig. 2C), with Gammaproteobacteria accoun- Group F4 was mainly composed of OTUs assigned to
ting for over 65% of total Proteobacteria sequences, fol- Ascomycota of the genus Cladosporium from different
lowed by Alphaproteobacteria, Deltaproteobacteria and environments (Fig. 3B; Table 4). In this community,
Betaproteobacteria (Fig. 2D). The bacterial community in Ascomycota represented more than 93% of the total
surface waters collected in Baker Channel in August (W5- fungal sequences (Fig. 3C).
3m) showed a similar pattern to those observed in bottom Samples collected in surface waters of the area close
waters grouped in B3. to the front of the glacier and in surface sediments from
Sediments from the area close to the fjord mouth the outside basin in August and surface waters of Baker
(W3-sed) and surface waters from the area close to the Channel in April were not included in any cluster
glacier (W1-0m) collected in August were not included in (Fig. 3A). These fungal communities were mainly com-
any main cluster (Fig. 2A). In these communities, OTUs posed of OTUs related to unclassified Fungi (W1-0m),
matching clones of Alphaproteobacteria, unclassified OTUs matching Chitridiomycota and Ascomycota (A5-
Bacteria and Gammaproteobacteria were representatives 3m), and OTUs assigned to Basidiomycota (W4-sed;
(Fig. 2B, Table 4) and a predominance of sequences of Fig. 3C).
Alphaproteobacteria and Gammaproteobacteria was
observed (Fig. 2D).
Composition of archaeal communities

With the exception of sample A4-3m, no sequences or


Composition of fungal communities
only low numbers (0–378 reads) were obtained for
Cluster analysis of fungal community at the OTU level Archaea during autumn (Table 3). During winter,
showed four main groups with more than 60% similarity although we recovered a higher number of sequences in
(Fig. 3A). Group F1 was composed of surface water com- most of the sampling sites (8–7031 reads), they repre-
munity in autumn, F2 included a mix of surface and sented a small fraction of all OTUs (Table 3). We
bottom waters in winter, F3 includes shallow sediments attributed this result to the set of primers used, which
and 100 m waters from Baker Channel collected in winter, were probably not able to detect typical Archaea from
and F4 was composed of bottom waters collected in glacial fjords waters. Among the sequences recovered,
autumn (Fig. 3A). The clustering was consistent with the more than 90% were related to sequences of
pattern observed in the ordination plots derived from Thaumarchaeota, with main OTUs matching clones pre-
PCoA (Fig. S3). viously detected in fresh and marine waters, sediments
Main OTUs in cluster F1 (surface waters from autumn in and cold environments (Table 4). Most of the character-
the fjord basin) were assigned to Basidiomycota, Asco- istic archaeal OTUs were found in samples collected in
mycota and uncharacterized Fungi (Fig. 3B; Table 4). winter (Table 4).

© 2015 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 17, 3882–3897
Microbial community in meltwaters of Patagonian fjords 3889

Fig. 3. Dendrogram based on Bray–Curtis similarity of the fungal assemblages in surface and bottom waters from autumn and winter in the
fjord adjacent to Jorge Montt glacier (A). Heatmap of the main fungal OTUs (B), and relative abundance (percentage of total sequences) of
main fungal groups (C). F1, F2, F3 and F4 represent the fungal communities identified with more than 60% of similarity.

Discussion retreat in the Patagonian Ice fields (Rivera et al., 2012).


During austral autumn and winter, we found lower con-
The impact of glacier meltwater on fjord microbes was centrations of nitrate and phosphate and higher concen-
assessed by analysing Bacteria, Archaea and Fungi com- trations of silicic acid in surface compared with bottom
munities in the water column and sediments of the fjord waters. These chemical characteristics are considered
adjacent to Jorge Montt, a glacier experiencing the fastest indicative of the influence of glacial-derived waters in

© 2015 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 17, 3882–3897
3890 M. H. Gutiérrez, P. E. Galand, C. Moffat and S. Pantoja

Patagonian fjords (González et al., 2013). Moreover, the Sequences of Proteobacteria, Actinobacteria and
thicker and more extensive surface meltwater layer, Bacteriodetes were also abundant in surface meltwater
overall lower salinities, and higher stratification of the during autumn (Fig. 2), with most of main OTUs related to
water column in autumn compared with winter indicate sequences previously identified in cold and freshwater
that the fjord is subjected to a higher freshwater discharge environments (Table 4). These Bacteria phyla are fre-
from the glacier in autumn. This seasonal variability in quently identified in glacial environments such as
meltwater content sets the framework to evaluate poten- cryoconite holes (Edwards et al., 2011; Cameron et al.,
tial changes in microbial communities exposed to increas- 2012), glacier-fed streams (Wilhelm et al., 2013), and
ing glacial melting. The autumn conditions, reflecting a fjords and polar waters influenced by glacial meltwater
higher contribution of meltwater, favoured the occurrence (Zeng et al., 2009; 2013; Piquet et al., 2010; 2011).
of a microbial community adapted to cold and freshwater Among Proteobacteria, Alphaproteobacteria, Betapro-
in the surface layer of the fjord. It indicates that a sus- teobacteria and Gammaproteobacteria showed similar
tained input of glacial-derived waters have the potential to relative abundance. This is consistent with previous
change microbial community composition in the findings showing high abundance of clones of Alphapro-
Patagonian fjords, displacing autochthonous microorgan- teobacteria, Betaproteobacteria and Gammaproteo-
isms and supporting the development of others, better bacteria in waters of Arctic fjords (Piquet et al., 2010)
adapted to live under the extreme conditions observed and with a predominance of Alphaproteobacteria and
under an ice melting scenario. Gammaproteobacteria in sea-ice environments (Brown
Bacterial richness was higher in autumn than winter in and Bowman, 2001) and in surface waters of high latitude
the surface layer of the fjord, suggesting both that melt- and fjords ecosystems influenced by glacial meltwaters
water supports the development of different bacterial taxa (Zeng et al., 2009; Piquet et al., 2011). In addition,
and that glacial-derived waters carry microorganisms to Alphaproteobacteria and Betaproteobacteria are common
the proglacial fjord. Although we lack direct measure- in cryoconite holes (Edwards et al., 2011, 2013, 2014;
ments of cell transport by meltwater, this would be Cameron et al., 2012) and Betaproteobacteria in
consistent with recent reports of substantial cellular subglacial Antarctic lakes (Christner et al., 2014). In par-
flux to downstream ecosystems from glacier surfaces ticular, the occurrence of abundant OTUs identified as
(Irvine-Fynn and Edwards, 2014). Cyanobacteria, identi- Comamonadaceae is consistent with previous observa-
fied previously as Synechococcus in cold and freshwater tions of this group of Betaproteobacteria in cryoconites
environments, were the predominant bacterial OTUs of holes (Cameron et al., 2012), glacial-fed streams
surface meltwater during austral autumn. Cyanobacteria (Wilhelm et al., 2013) and subglacial waters (Chen and
are considered the dominant phototrophic microorgan- Foght, 2007).
isms in cold freshwater environments (Vincent, 2000; Among Fungi, a predominance of Ascomycota over
Zakhia et al., 2008; Vincent and Quesada, 2012) and Basidiomycota-related sequences was observed in melt-
important contributors to nitrogen fixation in polar waters water in the fjord basin in autumn. Consistently, a sig-
(Vincent, 2000). In particular, the group Synechococcus is nificant proportion of Ascomycota have been isolated
a dominant part of picophytoplanktonic prokaryotes in from cryoconite sediments (Edwards et al., 2013), and in
freshwater environments (Callieri, 2007), and a conspicu- soils influenced by alpine glaciers, they dominate the
ous microorganism of Arctic and Antarctic water bodies fungal community in recently retreated soils (Zumsteg
(Vincent, 2000; Vincent et al., 2000), where potential et al., 2012). A different pattern was observed in surface
adaptations have been proposed for distinctive lineages waters of Baker Channel where a predominance of
(Huang et al., 2012). Cyanobacteria can play a main role sequences related to Chytridiomycetes was observed
in photosynthetic production of organic matter under low (Fig. 3C). Chytrid Fungi are ubiquitous in aquatic envi-
nutrient (Callieri and Stockner, 2002; Callieri, 2007) and ronments, where they are saprotrophs of plant material
oligotrophic conditions (Bell and Kalff, 2001; Callieri, and parasites of algae (James et al., 2006). This group
2007), in high latitude freshwaters bodies (Vincent and has been identified previously in periglacial soils
Quesada, 2012), and in glacial fjords (Piquet et al., 2014). (Freeman et al., 2009) as well as in cryoconites holes
Considering the low nitrate and phosphate concentrations (Edwards et al., 2013). Due to their role as saprotrophs,
observed in the surface oligotrophic meltwater of the fjord, we suggest that chytrids can play an important function
Cyanobacteria could contribute significantly to photosyn- in plant litter degradation in waters of Baker Channel,
thetic production in the waters adjacent to Jorge Montt where detritus from the adjacent forest may represent a
glacier during autumn. In agreement with that, Piquet significant source of organic matter. Our results show a
et al. (2014) found that meltwater inflow favoured the recurrent fraction of unclassified fungal sequences in
occurrence of nano- and picophytoplankton cells, includ- surface waters during autumn (up to 18% of total fungal
ing cyanophytes in waters of the Arctic fjords. sequences), suggesting that glacial waters, among

© 2015 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 17, 3882–3897
Microbial community in meltwaters of Patagonian fjords 3891

others, can be a source of unknown Fungi in Patagonian coastal waters, as well as in Arctic soil (Table 4). Although
fjords. Consistently, we observed higher abundance of the genus Cladosporium is a cosmopolitan Ascomycota
fungal OTUs in surface compared with bottom waters Fungi in terrestrial and air environments (Bensch et al.,
and inside the fjord compared with Baker Channel in 2012), it is also found in isolates from coastal and oceanic
autumn, as well as higher biomass of filamentous Fungi waters (Wang et al., 2012), from endosymbiotic commu-
in surface compared with bottom water and in meltwater nities of seaweed (Suryanarayanan, 2012) and from
close to the glacier front. We also found unclassified extreme environments such as the Dead Sea (Oren
fungal OTUs in surface sediments, as well as in the and Gunde-Cimerman, 2012) and salterns (Zajc et al.,
surface waters closest to the glacier front in winter, 2012).
where represented more than 80% of the total A different microbial assemblage was observed during
sequences. Our findings, together with the report by austral winter. The composition of this community was
Edwards and colleagues (2013) of abundant uncultured similar to those generally describing oceanic prokaryotes
Fungi in Arctic cryoconite holes, suggest that glacial (Biers et al., 2009; Zinger et al., 2011). The main OTUs
environments can be a significant reservoir of fungal included sequences related to Alphaproteobacteria,
diversity and biomass. Betaproteobacteria and Bacteriodetes identified previ-
A higher richness of bacterial OTUs was observed in ously in coastal and estuarine environments, Antarctic
bottom than in surface waters of the fjord basin during and Subantarctic waters, Pacific Arctic waters and sea-
autumn. Physicochemical characteristics of bottom water from the Baltic Sea (Table 4). During the same
waters are representative of marine conditions, suggest- period, sequences of Archaea were also recurrent in
ing that the bacterial community is more diverse in marine waters and sediments, with main OTUs related to
than meltwaters in the fjord. In waters deeper than 200 m sequences of Thaumarchaeota previously detected in
of the fjord basin during autumn, a distinctive microbial marine environments, including sediments of the Antarctic
community was detected. This community was character- region (Table 4). We suggest that a lower dilution of
ized by a dominance of Ascomycota Fungi and marine conditions, produced by a decrease in the melt-
Proteobacteria sequences, with Gammaproteobacteria water discharge from the glacier, results in the predomi-
accounting for more than 65% of total Proteobacteria. The nance of marine prokaryotes during winter. In addition, the
main bacterial OTUs in these bottom waters were related occurrence of mixing events breaking water column strati-
to sequences previously identified in deep ocean environ- fication could explain the presence of common microbial
ments, such as ocean crust, deep-sea mud volcanoes taxa in surface and bottom waters. Similarly, Piquet and
and hydrothermal vents, as well as subtidal sediments, colleagues (2011) found that in coastal Antarctic waters
deep-water coral reefs, and Subarctic and Antarctic wind-forced mixing events breaking meltwater stratifica-
waters (Table 4), suggesting the transport of microorgan- tion can produce changes in the composition of bacterial
isms in deep waters from the open ocean. That is con- community.
sistent with the model of circulation proposed for The Fungi community during winter was dominated by
Patagonian fjords, which describes the inflow of Subant- Basidiomycota-related sequences and was consistent
arctic waters to the deepest strata of the fjords (Silva with the global pattern described for continental airborne
et al., 1998). This deep inflow of waters of oceanic origin Fungi (Fröhlich-Nowoisky et al., 2012). Since Basidio-
was also observed in the proglacial fjord studied here mycota are predominant in terrestrial environments
(Moffat, 2014). (Buée et al., 2009), we suggest that air transport of
The composition of major groups of Fungi detected in allochthonous Fungi can play an important role in deter-
bottom waters during autumn (Fig. 3C) is consistent with mining fungal composition in Patagonian glaciers and
the global spatial distribution of major phyla of Fungi fjords.
observed in coastal and marine airborne environments Overall, the proglacial fjord surface meltwater commu-
(Fröhlich-Nowoisky et al., 2012), which agrees with the nity during autumn was characterized by a predominance
model of water circulation of these fjords and suggests of Cyanobacteria and Ascomycota over Basidiomycota-
that Subantarctic waters carry microbes to deep waters. A related sequences. In contrast, waters from winter were
strong dominance of sequences related to Ascomycota characterized by a higher proportion of Proteobacteria, a
was observed in the community of bottom waters during predominance of Basidiomycota sequences (over 80% of
autumn. This dominance appears to be a general charac- total fungal sequences in surface waters) and the occur-
teristic of fungal community in aquatic environments and rence of archaeal OTUs. In bottom waters from autumn,
sediments (Shearer et al., 2007; Mohamed and Martiny, Proteobacteria dominated bacterial community (more
2011). Main fungal OTUs in this community included than 70%), with a significant proportion of Gamma-
sequences related to Cladosporium spp. and unchar- proteobacteria, and the fungal community was dominated
acterized Fungi detected previously in freshwater and by Ascomycota.

© 2015 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 17, 3882–3897
3892 M. H. Gutiérrez, P. E. Galand, C. Moffat and S. Pantoja

In conclusion, our results suggest that meltwater from DNA extraction and 454 pyrosequencing
Jorge Montt glacier sets favourable conditions for the
DNA from filters and sediment samples (0.2 g) were
development of microbes adapted to cold and low salinity extracted and cleaned using a PowerSoil® DNA Isolation Kit
surface waters of the proglacial fjord during a high dis- and Power Clean® DNA Clean-up Kit (MOBIO Laboratories).
charge scenario. During these periods, Cyanobacteria Bacteria and Archaea 16S rRNA genes were amplified using
dominate the bacterial community and are likely involved primer sets 28F (5′-GAG TTT GAT CNT GGC TCA G-3′) and
in the photosynthetic synthesis of organic matter. Among 519R (5′-GTN TTA CNG CGG CKG CTG-3′), and 341F (5′-
Fungi, the presence of a significant fraction of airborne GYG CAS CAG KCG MGA AW-3′) and 958R (5′-GGA CTA
CVS GGG TAT CTA AT-3′) respectively. For Fungi, the inter-
and terrestrial taxa suggests the transport of
nal transcribed spacer (ITS) region was amplified by using
allochthonous Fungi by meltwater to the fjord, including the ITS1F (5′-CTT GGT CAT TTA GAG GAA GTA A-3′) and
unknown taxa. ITS4R (5′-TCC TCC GCT TAT TGA TAT GC-3′) set of
We also found that the deep-water circulation supplies primers. Amplification and sequencing on a 454 GS-FLX
marine microorganisms to the bottom waters of the fjord. system (Roche) with titanium chemistry were conducted in a
During autumn, this marine signal is diluted by meltwater, commercial laboratory (Research and Testing Laboratory,
but in winter a reduction in the thickness of the meltwater Lubbock, TX, USA) (Dowd et al., 2008).
layer and probably the occurrence of mixing events pro-
motes a dominance of marine microbes and sustain more
Sequence data processing
homogenous microbial communities in the water column.
In the relatively higher meltwater scenario, a fresh and The full 454-pyrosequencing data set is available at the
cold water signal was detected c. 20 km away from the National Center for Biotechnology Information Sequence
front of the glacier, suggesting that microbes associated Read Archive under the Accession No. SRR1607287 and
SRR160331. Raw sequences were analysed using the
with meltwater can represent a significant proportion of
Pyrotagger pipeline (Kunin and Hugenholtz, 2010). The reads
the microbial community in Patagonian fjords.
were first filtered by removing low-quality reads, then trimmed
to remove reads having ≥ 3% of bases with Phred values <27
(0.2% per-base error probability). This is recommended to
Experimental procedures
ensure that when clustering at 97%, the influence of errone-
Study area and sampling ous reads is minimized (Huse et al., 2010). The sequence
length threshold was set to 350 bp for Bacteria and Fungi and
Sample collection was carried out in the proglacial fjord adja- 200 bp for Archaea, and Uclust was used to cluster OTUs at
cent to Jorge Montt glacier, located in the northern section of a 97% similarity threshold. Representative sequences from
the Southern Patagonian Ice Field (48°20’S; 73°30’W; each OTU were classified by comparison with the
Fig. 1A). Field campaigns were conducted during austral Greengenes database (De Santis et al., 2006) for Archaea
autumn (25 April–2 May 2012) and winter (28–30 August and Bacteria and with SILVA (Quast et al., 2013) for Fungi.
2012) onboard the vessel Don Tito. Water was collected with Pyrotagger outputs, including OTU distribution among
Niskin bottles from surface (0–3 m) and bottom waters (50– samples and taxonomic affiliation, were used as input for
300 m) in sites along a 20 km gradient, which includes waters analyses in the qiime software package version 1.6.0
from the area close to the glacier terminus and from Baker (Caporaso et al., 2010) after removal of potential chimeras.
channel (outside the fjord basin; Fig. 1, Table 1). One litre of Analysis of similarity in community composition and diversity
pre-sieved seawater (210 μm) was filtered through 0.45 μm (Chao1) was carried out after resampling with the rarefaction
sterile membrane filters (Millipore) and stored at −20°C for method using the minimum number of sequences per sample
DNA analyses. Sediments were collected with a dredge at for Bacteria and Fungi and a subsample of 60 sequences for
two sites, inside (38 m depth) and outside (84 m depth) the Archaea. Similarity was estimated at the OTU level based on
fjord basin during the winter campaign (Fig. 1A, Table 1). the Bray–Curtis distance matrix index and hierarchical cluster
Hydrographic data were recorded using a SeaBird 25 analysis (UPGMA clustering). A PCoA was carried out in R
CTD in April 2012 and a RBR XRX-620 CTD in August version 3.1.2 using the package vegan (Oksanen et al.,
2012. 2013) to verify the clustering pattern obtained with UPGMA
Samples for quantification of particulate chlorophyll and clustering.
dissolved inorganic nutrients were collected from c. 1 L sea- OTU heatmaps were produced to identify the contribution
water filtered throughout 0.7 µm glass fiber filters. of individual OTUs to the community composition of the dif-
Chlorophyll-α was analysed by fluorometry (Parsons et al., ferent samples. OTU heatmaps were filtered to display the
1984) and nutrients by spectrophotometry (Strickland and representative OTUs defined as the ones containing more
Parsons, 1972). Prokaryotes and fungal filaments abundance than 150 or 60 sequences per OTUs for Fungi and Bacteria
were analysed by epifluorescence microscopy using 4,6- respectively. Rarefaction curves for each microbial commu-
diamidino-2-phenylindole (Porter and Feig, 1980) and nity were generated from the means of 100 randomized data
Calcofluor White (Gutiérrez et al., 2011) from 10 to 50 ml of sets using the software EstimateS version 9 (Colwell, 2013).
water subsamples collected directly from the Niskin bottle Two-way analysis of variance was applied to test statistical
and fixed with 2% and 3% formaldehyde (final concentration) differences between periods and depths on the rank-
respectively. transformed data of environmental variables and community

© 2015 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 17, 3882–3897
Microbial community in meltwaters of Patagonian fjords 3893
diversity. This transformation allows dealing with data that do Bensch, K., Braun, U., Groenewald, J.Z., and Crous, P.W.
not meet the assumptions of normality (Conover and Iman, (2012) The genus Cladosporium. Stud Mycol 15: 1–401.
1981; Neter et al., 1996). Biers, E.J., Sun, S., and Howard, E.C. (2009) Prokaryotic
genomes and diversity in surface ocean waters: interrogat-
ing the global ocean sampling metagenome. Appl Environ
Acknowledgements Microbiol 75: 2221–2229.
Boldt, K.V., Nittrouer, C., Hallet, B., Koppes, M.N., Forrest,
This research was funded by FONDECYT Grants Nos.
B.K., Wellner, J.S., and Anderson, J.B. (2013) Modern
11110515 and 791220026, CONICYT, Chile. A research visit
rates of glacial sediment accumulation along a 15° S-N
of M. H. Gutiérrez to the Observatoire Océanologique de
transect in fjords from the Antarctic peninsula to southern
Banyuls was funded by the International Associated Labora-
Chile. J Geophys Res Earth Surf 118: 2072–2088.
tory MORFUN. C. Moffat and the collection and analysis of
Brown, M.V., and Bowman, J.P. (2001) A molecular
hydrographic data was funded by FONDECYT Grant No.
phylogenetic survey of sea-ice microbial communities
11100362. P. Galand was supported by the Agence Nationale
(SIMCO). FEMS Microbiol Ecol 35: 267–275.
de la Recherche (ANR) project MICADO (ANR-11JSV7-003-
Buée, M., Reich, M., Murat, C., Morin, E., Nilsson, R.H., Uroz,
01). We acknowledge C. Córdova, A. M. Jara, R. Mansilla
S., and Martin, F. (2009) 454 Pyrosequencing analyses of
and C. Iturra for their help during fieldwork in Patagonia and
forest soils reveal an unexpectedly high fungal diversity.
laboratory analysis, R. Veas for his assistance with statistical
New Phytol 184: 449–456.
analysis and Waters of Patagonia and the crew of the vessel
Callieri, C. (2007) Picophytoplankton in freshwater ecosys-
Don Tito for their support during fieldwork. Partial funding
tems: the importance of small-sized phototrophs. Freshw
was obtained from Program COPAS Sur-Austral PFB-31. We
Rev 1: 1–28.
are grateful to the Hanse-Wissenschaftskolleg (HWK),
Callieri, C., and Stockner, J.G. (2002) Freshwater autotrophic
Delmenhorst, Germany, for the Fellowship awarded to S.
phytoplankton: a review. J Limnol 61: 1–14.
Pantoja.
Cameron, K.A., Hodson, A.J., and Osborn, A.M. (2012)
Structure and diversity of bacterial, eukaryotic and
References archaeal communities in glacial cryoconite holes from
the Arctic and the Antarctic. FEMS Microbiol Ecol 82: 254–
Acosta-Gonzalez, A., Rossello-Mora, R., and Marques, S. 267.
(2013) Characterization of the anaerobic microbial commu- Caporaso, J.G., Kuczynski, J., Stombaugh, J., Bittinger, K.,
nity in oil-polluted subtidal sediments: aromatic biodegra- Bushman, F.D., Costello, E.K., et al. (2010) QIIME allows
dation potential after the Prestige oil spill. Environ Microbiol analysis of high-throughput community sequencing data.
15: 77–92. Nat Methods 7: 335–336.
Allers, E., Wright, J.J., Konwar, K.M., Howes, C.G., Beneze, Chen, S.M., and Foght, J.M. (2007) Cultivation-independent
E., Hallam, S.J., and Sullivan, M.B. (2013) Diversity and and -dependent characterization of Bacteria resident
population structure of Marine Group A bacteria in the beneath John Evans glacier. FEMS Microbiol Ecol 59:
Northeast subarctic Pacific Ocean. ISME J 7: 256–268. 318–330.
Alonso, L.M., Kleiner, D., and Ortega, E. (2008) Spores of the Christner, B.C., Priscu, J.C., Achberger, A.M., Barbante, C.,
mycorrhizal fungus Glomus mosseae host yeasts that Carter, S.P., Christianson, K., et al. (2014) A microbial eco-
solubilize phosphate and accumulate polyphosphates. system beneath the West Antarctic ice sheet. Nature 512:
Mycorrhiza 18: 197–204. 310–313.
Arenz, B.E., Held, B.W., Jurgens, J.A., Farrell, R.L., and Clingenpeel, S., Macur, R.E., Kan, J., Inskeep, W.P., Lovalvo,
Blanchette, R.A. (2006) Fungal diversity in soils and his- D., Varley, J., et al. (2011) Yellowstone Lake: high-energy
toric wood from the Ross Sea Region of Antarctica. Soil geochemistry and rich bacterial diversity. Environ Microbiol
Biol Biochem 38: 3057–3064. 13: 2172–2185.
Arhipova, N., Gaitnieks, T., Donis, J., Stenlid, J., and Vasaitis, Colwell, R.K. (2013) EstimateS, version 8.2: statistical esti-
R. (2011) Decay, yield loss and associated fungi in stands mation of species richness and shared species from
of grey alder (Alnus incana) in Latvia. Forestry 84: 337– samples. Version 9 [WWW document]. URL http://
348. www.purl.oclc.org/estimates.
Auguet, J.-C., and Casamayor, E.O. (2012) Partitioning of Conover, W.J., and Iman, R.L. (1981) Rank transformation as
Thaumarchaeota populations along environmental gradi- a bridge between parametric and non parametric statistics.
ents in high mountain lakes. FEMS Microbiol Ecol 84: Am Stat 35: 124–129.
154–164. Cordier, T., Robin, C., Capdevielle, X., Desprez-Loustau,
Bell, T., and Kalff, J. (2001) The contribution of M.L., and Vacher, C. (2012) Spatial variability of
picophytoplankton in marine and freshwater systems of phyllosphere fungal assemblages: genetic distance pre-
different trophic status and depth. Limnol Oceanogr 46: dominates over geographic distance in a European beech
1243–1248. stand (Fagus sylvatica). Fungal Ecol 5: 509–520.
Bensch, K., Groenewald, J.Z., Dijksterhuis, J., Starink- Cowie, R.O.M., Maas, E.W., and Ryan, K.G. (2011) Archaeal
Willemse, M., Andersen, B., Summerell, B.A., et al. (2010) diversity revealed in Antarctic sea ice. Antarct Sci 23: 531–
Species and ecological diversity within the Cladosporium 536.
cladosporioides complex (Davidiellaceae, Capnodiales). Crosbie, N.D., Pöckl, M., and Weisse, T. (2003a)
Stud Mycol 67: 1–94. Dispersal and phylogenetic diversity of non marine

© 2015 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 17, 3882–3897
3894 M. H. Gutiérrez, P. E. Galand, C. Moffat and S. Pantoja
picocyanobacteria, inferred from 16S rRNA gene and of archaeal assemblages in the coastal NW Mediterranean
cpcBA-intergenic spacer sequence analyses. Appl Environ Sea (Blanes Bay Microbial Observatory). Limnol Oceanogr
Microbiol 69: 5716–5721. 55: 2117–2125.
Crosbie, N.D., Pöckl, M., and Weisse, T. (2003b) Rapid Gillan, D.C., and Danis, B. (2007) The archaebacterial com-
establishment of clonal isolates of freshwater autotrophic munities in Antarctic bathypelagic sediments. Deep Sea
picoplankton by single-cell and single-colony sorting. J Res Part II Top Stud Oceanogr 54: 1682–1690.
Microbiol Methods 55: 361–370. Glatz, R.E., Lepp, P.W., Ward, B.B., and Francis, C.A. (2006)
Croué, J., West, N.J., Escande, M.-L., Intertaglia, L., Planktonic microbial community composition across steep
Lebaron, P., and Suzuki, M.T. (2013) A single physical/chemical gradients in permanently ice-covered
betaproteobacterium dominates the microbial community Lake Bonney, Antarctica. Geobiology 4: 53–67.
of the crambescidine-containing sponge Crambe crambe. González, H.E., Castro, L., Daneri, G., Iriarte, J.L., Silva, N.,
Sci Rep 3: art2583. Vargas, C., et al. (2011) Seasonal plankton variability in
Dávila, P., Figueroa, D., and Muller, E. (2002) Freshwater Chilean Patagonia fjords: carbon flow through the pelagic
input into the coastal ocean and its relation with the salinity food web of the Aysen Fjord and plankton dynamics in
distribution off austral Chile (35–55°S). Cont Shelf Res 22: the Moraleda Channel basin. Cont Shelf Res 31: 225–
521–534. 243.
De Santis, T.Z., Hugenholtz, P., Larsen, N., Rojas, M.E., González, H.E., Castro, L.R., Daneri, G., Iriarte, J.L., Silva,
Brodie, L., Keller, K., et al. (2006) Greengenes, a Chimera- N., Tapia, F., et al. (2013) Land–ocean gradient in haline
checked 16S rRNA gene database and workbench stratification and its effects on plankton dynamics and
compatible with ARB. Appl Environ Microbiol 72: 5069– trophic carbon fluxes in Chilean Patagonian fjords (47–
5072. 50°S). Prog Oceanogr 119: 32–47.
Dowd, S.E., Callaway, T.R., Wolcott, R.D., Sun, Y., Gunde-Cimerman, N., Sonjak, S., Zalar, P., Frisvad, J.C.,
McKeehan, T., Hagevoort, R.G., and Edrington, T.S. (2008) Diderichsen, B., and Plemenitas, A. (2003) Extremophilic
Evaluation of the bacterial diversity in the feces of cattle fungi in Arctic ice: a relationship between adaptation to low
using 16S rDNA bacterial tag-encoded FLX amplicon temperature and water activity. Phys Chem Earth (2002)
pyrosequencing (bTEFAP). BMC Microbiol 8: 125. 28: 1273–1278.
Edwards, A., Anesio, A.M., Rassner, S.M., Sattler, B., Gutiérrez, M.H., Pantoja, S., Tejos, E., and Quiñones, R.A.
Hubbard, B., Perkins, W.T., et al. (2011) Possible interac- (2011) The role of fungi in processing marine organic
tions between bacterial diversity, microbial activity and matter in the upwelling ecosystem off Chile. Mar Biol 158:
supraglacial hydrology of cryoconite holes in Svalbard. 205–219.
ISME J 5: 150–160. Hong, J.W., Fomina, M., and Gadd, G.M. (2010) F-RISA
Edwards, A., Douglas, B., Anesio, A.M., Rassner, A.M., fungal clones as potential bioindicators of organic and
Tristram, D.L., Irvine-Fynn, T.D.L., et al. (2013) A distinctive metal contamination in soil. J Appl Microbiol 109: 415–
fungal community inhabiting cryoconite holes on glaciers in 430.
Svalbard. Fungal Ecol 6: 168–176. Hu, L., Cao, L., and Zhang, R. (2013) Bacterial and fungal
Edwards, A., Mur, L.A.J., Girdwood, S.E., Anesio, A.M., taxon changes in soil microbial community composition
Stibal, M., Rassner, S.M.E., et al. (2014) Coupled induced by short-term biochar amendment in red oxidized
cryoconite ecosystem structure–function relationships are loam soil. World J Microbiol Biotechnol 3: 1085–1092.
revealed by comparing bacterial communities in alpine and Huang, S., Wilhelm, S.W., Harvey, H.R., Taylor, K., Jiao, N.,
Arctic glaciers. FEMS Microbiol Ecol 89: 222–237. and Chen, F. (2012) Novel lineages of Prochlorococcus
Ernst, A., Becker, S., Wollenzien, U.I.A., and Postius, C. and Synechococcus in the global oceans. ISME J 6: 285–
(2003) Ecosystem-dependent adaptive radiations of 297.
picocyanobacteria inferred from 16S rRNA and ITS-1 Huse, S.M., Welch, D.M., Morrison, H.G., and Sogin, M.L.
sequence analysis. Microbiology 149: 217–228. (2010) Ironing out the wrinkles in the rare biosphere
Flynn, T.M., Sanford, R.A., Ryu, H., Bethke, C.M., Levine, through improved OTU clustering. Environ Microbiol 12:
A.D., Ashbolt, N.J., et al. (2013) Functional microbial diver- 1889–1898.
sity explains groundwater chemistry in a pristine aquifer. Irvine-Fynn, T.D.L., and Edwards, A. (2014) A frozen asset:
BMC Microbiol 13: 146. the potential of flow cytometry in constraining the glacial
Freeman, K.R., Martin, A.P., Karki, D., Lynch, R.C., Mitter, biome. Cytometry A 85A: 3–7.
M.S., Meyer, A.F., et al. (2009) Evidence that chytrids James, T.Y., Letcher, P.M., Longcore, J.E., Mozley-
dominate fungal communities in high-elevation soils. Proc Standridge, S.E., Porter, D., Powell, M.J., et al. (2006)
Natl Acad Sci USA 106: 18315–18320. A molecular phylogeny of the flagellated fungi
Fröhlich-Nowoisky, J., Pickersgilla, D.A., Desprésa, V.R., and (Chytridiomycota) and description of a new phylum
Pöschla, U. (2009) High diversity of fungi in air particulate (Blastocladiomycota). Mycologia 98: 860–871.
matter. Proc Natl Acad Sci USA 106: 12814–12819. Jensen, S., Neufeld, J.D., Birkeland, N.-K., Hovland, M., and
Fröhlich-Nowoisky, J., Burrows, S.M., Xie, Z., Engling, G., Murrell, J.C. (2008) Insight into the microbial community
Solomon, P.A., Fraser, M.P., et al. (2012) Biogeography in structure of a Norwegian deep-water coral reef environ-
the air: fungal diversity over land and oceans. ment. Deep Sea Res Part I Oceanogr Res Pap 55: 1554–
Biogeosciences 9: 1125–1136. 1563.
Galand, P.E., Gutiérrez-Provecho, C., Massana, R., Gasol, Kato, S., Chan, C., Itoh, T., and Ohkuma, M. (2013) Func-
J.M., and Casamayor, E.O. (2010) Inter-annual recurrence tional gene analysis of freshwater iron-rich flocs at

© 2015 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 17, 3882–3897
Microbial community in meltwaters of Patagonian fjords 3895
circumneutral pH and isolation of a stalk-forming ture in the vicinity of Antarctic icebergs. Deep Sea Res Part
microaerophilic iron-oxidizing bacterium. Appl Environ II Top Stud Oceanogr 58: 1407–1421.
Microbiol 79: 5283–5290. Neter, J., Kutner, M., Wasserman, W., and Nashtsheim, C.
Kunin, V., and Hugenholtz, P. (2010) PyroTagger: a fast, (1996) Applied Linear Statistical Models, 4th edn. Chicago,
accurate pipeline for analysis of rRNA amplicon IL, USA: McGraw–Hill/Irwin.
pyrosequence data. Open J 1: 1–8. Newbold, L.K., Oliver, A.E., Booth, T., Tiwari, B., De Santis,
La Cono, V., La Spada, G., Arcadi, E., Placenti, F., Smedile, T., Maguire, M., et al. (2012) The response of marine
F., Ruggeri, G., et al. (2013) Partaking of Archaea to picoplankton to ocean acidification. Environ Microbiol 14:
biogeochemical cycling in oxygen-deficient zones of 2293–2307.
meromictic saline Lake Faro (Messina, Italy). Environ Oksanen, A.J., Blanchet, F.G., Kindt, R., Minchin, P.R., Hara,
Microbiol 15: 1717–1733. R.B.O., Simpson, G.L., et al. (2013) vegan: community
Li, H., Yu, Y., Luo, W., Zeng, Y., and Chen, B. (2009) Bacterial Ecology Package. R package version 2.0–5.
diversity in surface sediments from the Pacific Arctic Oliveira, B.R., Barreto Crespo, M.T., San Roma, M.V.,
Ocean. Extremophiles 13: 233–246. Benoliel, M.J., Samson, R.A., and Pereira, V.J. (2013) New
Liao, P.C., Huang, B.H., and Huang, S. (2007) Microbial insights concerning the occurrence of fungi in water
community composition of the Danshui river estuary of sources and their potential pathogenicity. Water Res 47:
Northern Taiwan and the practicality of the phylogenetic 6338–6347.
method in microbial barcoding. Microb Ecol 54: 497– Oren, A., and Gunde-Cimerman, N. (2012) Fungal life in the
507. Dead Sea. In Biology of Marine Fungi. Raghukumar, C.
Liu, Y., Tandong, Y., Jiao, N., Tian, L., Hu, A., Yu, W., and Li, (ed.). Berlin Heidelberg, Germany: Springer-Verlag, pp.
S. (2011) Microbial diversity in the snow, a moraine lake 115–132.
and a stream in Himalayan glacier. Extremophiles 15: 411– Pachiadaki, M.G., Kallionaki, A., Dählmann, A., De Lange,
421. G.J., and Kormas, K.A. (2011) Diversity and spatial distri-
Liu, Z., Huang, S., Sun, G., Xu, Z., and Xu, M. (2012) bution of prokaryotic communities along a sediment verti-
Phylogenetic diversity, composition and distribution of cal profile of a deep-sea mud volcano. Microb Ecol 62:
bacterioplankton community in the Dongjiang River, China. 655–668.
FEMS Microbiol Ecol 80: 30–44. Parsons, T.R., Maita, Y., and Lalli, C.M. (1984) A Manual of
Macia-Vicente, J.G., Ferraro, V., Burruano, S., and Chemical and Biological Methods for Seawater Analysis.
Lopez-Llorca, L.V. (2012) Fungal assemblages associated Oxford: Pergamon Press.
with roots of halophytic and non-halophytic plant species Pinhassi, J., Zweifel, U.L., and Hagström, A. (1997) Domi-
vary differentially along a salinity gradient. Microb Ecol 64: nant marine bacterioplankton species found among
668–679. colony-forming bacteria. Appl Environ Microbiol 63: 3359–
Mansergh, S., and Zehr, J.P. (2014) Vibrio diversity and 3366.
dynamics in the Monterey Bay upwelling region. Front Piquet, A.M.-T., Scheepens, J.F., Bolhuis, H., Wiencke, C.,
Microbiol 5: art48. and Buma, A.G.J. (2010) Variability of protistan and bac-
Margesin, R., and Miteva, V. (2011) Diversity and ecology of terial communities in two Arctic fjords (Spitsbergen). Polar
psychrophilic microorganisms. Res Microbiol 162: 346– Biol 33: 1521–1536.
361. Piquet, A.M.-T., Bolhuis, H., Meredith, M.P., and Buma,
Moffat, C. (2014) Wind-driven modulation of warm water A.G.J. (2011) Shifts in coastal Antarctic marine microbial
supply to a proglacial fjord, Jorge Montt glacier, Patagonia. communities during and after meltwater-related surface
Geophys Res Lett 41: 3943–3950. stratification. FEMS Microbiol Ecol 76: 413–427.
Mohamed, D.J., and Martiny, J.B.H. (2011) Patterns of fungal Piquet, A.M.-T., van de Poll, W.H., Visser, R.J.W., Wiencke,
diversity and composition along a salinity gradient. ISME J C., Bolhuis, H., and Buma, A.G.J. (2014) Springtime
5: 379–388. phytoplankton dynamics in Arctic Krossfjorden and
Monchy, S., Sanciu, G., Jobard, M., Rasconi, S., Kongsfjorden (Spitsbergen) as a function of glacier prox-
Gerphagnon, M., Chabé, M., et al. (2011) Exploring and imity. Biogeosciences 11: 2263–2279.
quantifying fungal diversity in freshwater lake ecosy- Porter, K.G., and Feig, Y.S. (1980) The use of DAPI for
stems using rDNA cloning/sequencing and SSU tag identifying and counting aquatic microflora. Limnol
pyrosequencing. Environ Microbiol 13: 1433–1453. Oceanogr 25: 943–948.
Montero, P., Daneri, G., González, H.E., Iriarte, J.L., Tapia, Prabagaran, S.R., Manorama, R., Delille, D., and Shivaji, S.
F.J., Lizárraga, L., et al. (2011) Seasonal variability of (2007) Predominance of Roseobacter, Sulfitobacter,
primary production in a fjord ecosystem of the Chilean Glaciecola and Psychrobacter in seawater collected off
Patagonia: implications for the transfer of carbon within Ushuaia, Argentina, Sub-Antarctica. FEMS Microbiol Ecol
pelagic food webs. Cont Shelf Res 31: 202–215. 59: 342–355.
Møller, A.K., Søborg, D.A., Al-Soud, W.A., Sørensen, S.J., Quast, C., Pruesse, E., Yilmaz, P., Gerken, J., Schweer, T.,
and Kroer, N. (2013) Bacterial community structure in Yarza, P., et al. (2013) The SILVA ribosomal RNA gene
High-Arctic snow and freshwater as revealed by database project: improved data processing and web-
pyrosequencing of 16S rRNA genes and cultivation. Polar based tools. Nucleic Acids Res 41: 590–596.
Res 32: art17390. Ramírez-Cruz, V., Guzmán, G., Villalobos-Arámbula, A.R.,
Murray, A.E., Peng, V., Tyler, C., and Wagh, P. (2011) Marine Rodríguez, A., Matheny, P.B., Sánchez-García, M., and
bacterioplankton biomass, activity and community struc- Guzmán-Dávalos, L. (2013) Phylogenetic inference and

© 2015 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 17, 3882–3897
3896 M. H. Gutiérrez, P. E. Galand, C. Moffat and S. Pantoja
trait evolution of the psychedelic mushroom genus Strickland, J.D., and Parsons, T.R. (1972) A Manual of Sea
Psilocybe sensu lato (Agaricales). Botany 91: 573–591. Water Analysis. Bulletin No. 125. Otawa, Canada: Fisher-
Renker, C., Blanke, V., Borstler, B., Heinrichs, J., and Buscot, ies Research Board of Canada.
F. (2004) Diversity of Cryptococcus and Dioszegia yeasts Suryanarayanan, T.S. (2012) Fungal endosymbionts of sea-
(Basidiomycota) inhabiting arbuscular mycorrhizal roots or weeds. In Biology of Marine Fungi. Raghukumar, C. (ed.).
spores. FEMS Yeast Res 4: 597–603. Berlin Heidelberg, Germany: Springer-Verlag, pp. 53–69.
Riemann, L., Leitet, C., Pommier, T., Simu, K., Holmfeldt, K., Tang, C., Madigan, M.T., and Lanoil, B. (2013) Bacterial and
Larsson, U., and Hagström, A. (2008) The native archaeal diversity in sediments of west lake Bonney,
bacterioplankton community in the central Baltic sea is McMurdo dry valleys, Antarctica. Appl Environ Microbiol
influenced by freshwater bacterial species. Appl Environ 79: 1034–1038.
Microbiol 74: 503–515. Thorn, R.G., Reddy, C.A., Harris, D., and Paul, E.A. (1996)
Rignot, E., Rivera, A., and Casassa, G. (2003) Contribution of Isolation of Saprophytic Basidiomycetes from soil. Appl
the Patagonia Icefields of South America to global sea level Environ Microbiol 62: 4288–4292.
rise. Science 302: 434–437. Timling, I., Walker, D.A., Nusbaum, C., Lennon, N.J., and
Rivera, A., Benham, T., Casassa, G., Bamber, J., and Taylor, D.L. (2014) Rich and cold: diversity, distribution and
Dowdeswell, J.A. (2007) Ice elevation and areal changes drivers of fungal communities in patterned-ground ecosys-
of glaciers from the Northern Patagonia Icefield, Chile. tems of the North American Arctic. Mol Ecol 23: 3258–
Glob Planet Change 59: 126–137. 3272.
Rivera, A., Corripio, J., Bravo, C., and Cisternas, S. (2012) Uetake, J., Yoshimura, Y., Nagatsuka, N., and Kanda, H.
Glaciar Jorge Montt (Chilean Patagonia) dynamics derived (2012) Isolation of oligotrophic yeasts from supraglacial
from photos obtained by fixed cameras and satellite image environments of different altitude on the Gulkana glacier
feature tracking. Ann Glaciol 53: 147–155. (Alaska). FEMS Microbiol Ecol 82: 279–286.
Santelli, C.M., Orcutt, B.N., Banning, E., Bach, W., Moyer, Vargas, C.A., Martinez, R.A., San Martin, V., Aguayo, M.,
C.L., Sogin, M.L., et al. (2008) Abundance and diversity of Silva, N., and Torres, R. (2011) Allochthonous subsidies of
microbial life in ocean crust. Nature 453: 653–656. organic matter across a lake–river–fjord landscape in the
Schmidtova, J., Hallam, S.J., and Baldwin, S.A. (2009) Chilean Patagonia: implications for marine zooplankton in
Phylogenetic diversity of transition and anoxic zone bacte- inner fjord areas. Cont Shelf Res 31: 187–201.
rial communities within a near-shore anoxic basin: Nitinat Vincent, W.F. (2000) Cyanobacterial dominance in the polar
Lake. Environ Microbiol 11: 3233–3251. regions. In The Ecology of Cyanobacteria. Whitton, B.A.,
Scopel, M., dos Santos, O., Frasson, A.P., Abraham, W.-R., and Pottts, M. (eds). Dordrecht, Netherlands: Kluwer Aca-
Tasca, T., Henriques, A.T., and Macedo, A.J. (2013) Anti- demic Publisher, pp. 321–340.
Trichomonas vaginalis activity of marine-associated fungi Vincent, W.F., and Quesada, A. (2012) Cyanobacteria in high
from the South Brazilian Coast. Exp Parasitol 133: 211– latitude lakes, rivers and seas. In Ecology of Cyanobacteria
216. II: Their Diversity in Space and Time. Whitton, B.A. (ed.).
Sepúlveda, J., Pantoja, S., and Hughen, K.A. (2011) Sources New York, NY, USA: Springer-Verlag, pp. 371–385.
and distribution of organic matter in northern Patagonia Vincent, W.F., Bowman, J.P., Rankin, L.M., and McMeekin,
fjords, Chile (∼44–47°S): a multi-tracer approach for T.A. (2000) Phylogenetic diversity of picocyanobacteria in
carbon cycling assessment. Cont Shelf Res 31: 315–329. Arctic and Antarctic ecosystems. In Microbial Biosystems:
Shearer, C.A., Descals, E., Kohlmeyer, B., Kohlmeyer, J., New Frontiers. Proceedings of the 8th International Sym-
Marvanova, L., Padgett, D., et al. (2007) Fungal biodiver- posium on Microbial Ecology. Bell, R., Brylinsky, C.M., and
sity in aquatic habitats. Biodivers Conserv 16: 49–67. Johnson-Green, M. (eds). Halifax, Canada: Atlantic
Silva, N., Calvete, C., and Sievers, H. (1998) Masas de agua Canada Society for Microbial Ecology, pp. 317–322.
y circulación para algunos canales australes entre Puerto Wang, F., Zhou, H., Meng, J., Peng, X., Jiang, L., Sun, P.,
Montt y Laguna San Rafael, Chile (Crucero Cimar-Fiordo et al. (2009) GeoChip-based analysis of metabolic
1). Cienc Tecnol Mar 21: 17–48. diversity of microbial communities at the Juan de Fuca
Silva, N., Vargas, C.A., and Prego, R. (2011) Land–ocean Ridge hydrothermal vent. Proc Natl Acad Sci USA 106:
distribution of allochthonous organic matter in surface sedi- 4840–4845.
ments of the Chiloé and Aysén interior seas (Chilean North- Wang, G., Wang, X., Liu, X., and Li, Q. (2012) Diversity and
ern Patagonia). Cont Shelf Res 31: 330–339. biogeochemical function of planktonic fungi in the ocean. In
Singh, P., Raghukumar, K., Meena, R.M., Verma, P., and Biology of Marine Fungi. Raghukumar, C. (ed.). Berlin
Shouche, Y. (2012) Fungal diversity in deep-sea sediments Heidelberg, Germany: Springer-Verlag, pp. 71–88.
revealed by culture-dependent and culture-independent Wilhelm, L., Singer, G.A., Fasching, C., Battin, T.J., and
approaches. Fungal Ecol 5: 543–553. Besemer, K. (2013) Microbial biodiversity in glacier-fed
Sjöstedt, J., Martiny, J.B., Munk, P., and Riemann, L. (2014) streams. ISME J 7: 1651–1660.
Abundance of broad bacterial taxa in the sargasso sea Willis, M.J., Melkonian, A.K., Pritchard, M.E., and Rivera, A.
explained by environmental conditions but not water mass. (2012) Ice loss from the Southern Patagonian Ice Field,
Appl Environ Microbiol 80: 2786–2795. South America, between 2000 and 2012. Geophys Res
Stenström, E., Ndobe, N.E., Jonsson, M., Stenlid, J., and Lett 39: L17501. doi:10.1029/2012GL053136.
Menkis, A. (2013) Root-associated fungi of healthy-looking Wirsel, S.G.R., Leibinger, W., Ernst, M., and Mendgen, K.
Pinus sylvestris and Picea abies seedlings in Swedish (2001) Genetic diversity of fungi closely associated with
forest nurseries. Scand J For Res 29: 12–21. common reed. New Phytol 149: 589–598.

© 2015 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 17, 3882–3897
Microbial community in meltwaters of Patagonian fjords 3897
Xing, P., Hahn, M.W., and Wu, Q.L. (2009) Low taxon Zinger, L., Amaral-Zettler, L.A., Fuhrman, J.A., Horner-
richness of bacterioplankton in high-altitude lakes of Devine, M.C., Huse, S.M., Welch, D.B.M., et al. (2011)
the Eastern Tibetan Plateau, with a predominance of Global patterns of bacterial beta-diversity in seafloor and
bacteroidetes and Synechococcus spp. Appl Environ seawater ecosystems. PLoS ONE 6: 1–11.
Microbiol 22: 7017–7025. Zumsteg, A., Luster, J., Göransson, H., Smittenberg, R.H.,
Yurkov, A., Kruger, D., Begerow, D., Arnold, N., and Tarkka, Brunner, I., Bernasconi, S.M., et al. (2012) Bacterial,
M.T. (2012) Basidiomycetous yeasts from boletales fruiting Archaeal and fungal succession in the forefield of a reced-
bodies and their interactions with the mycoparasite ing glacier. Microb Ecol 63: 552–564.
Sepedonium chrysospermum and the host fungus Paxillus.
Microb Ecol 63: 295–303.
Zajc, J., Zalar, P., Plemenitaš, A., and Gunde-Cimerman, N.
(2012) The mycobiota of the salterns. In Biology of Marine Supporting information
Fungi. Raghukumar, C. (ed.). Berlin Heidelberg, Germany:
Additional Supporting Information may be found in the online
Springer-Verlag, pp. 133–158.
version of this article at the publisher’s web-site:
Zakhia, F., Jungblut, A.-D., Taton, A., Vincent, W.F., and
Wilmotte, A. (2008) Cyanobacteria in cold ecosystems. Fig. S1. Rarefaction curves for Bacteria (A), Fungi (B) and
In Psychrophiles: from Biodiversity to Biotechnology. Archaea (C) for each sampling sites and depths in the fjord
Margesin, R., Schinner, F., Marx, J.C., and Gerday, C. adjacent to the Jorge Montt glacier.
(eds). Berlin Heidelberg, Germany: Springer-Verlag, pp. Fig. S2. Principal coordinate analysis (PCoA) based on
121–135. Bray–Curtis distance matrix of OTUs of Bacteria in the fjord
Zeng, Y., Zheng, T., and Li, H. (2009) Community composi- adjacent to the Jorge Montt glacier. Colours in three-
tion of the marine bacterioplankton in Kongsfjorden dimensional plots represent the groups identified by the
(Spitsbergen) as revealed by 16S rRNA gene analysis. UPGMA clustering analysis. NC indicates samples not
Polar Biol 32: 1447–1460. grouped by the clustering method. PCo1, PCo2 and PCo3
Zeng, Y.-X., Zhang, F., He, J.-F., Lee, S.H., Qiao, Z.-Y., Yu, Y., explained 67% of the variability.
and Li, H.-R. (2013) Bacterioplankton community structure Fig. S3. Principal coordinate analysis (PCoA) on the basis of
in the Arctic waters as revealed by pyrosequencing of 16S Bray–Curtis distance matrix of OTUs of bacterial (A) and
rRNA genes. Antonie Van Leeuwenhoek 103: 1309–1319. fungal (B) communities in the fjord adjacent to the Jorge
Zhang, R., Liu, B., Lau, S.C., Ki, J.S., and Qian, P.Y. (2007) Montt glacier. Colours in three-dimensional plots represent
Particle-attached and free-living bacterial communities in a the groups identified by UPGMA clustering analysis. NG indi-
contrasting marine environment: Victoria Harbor, Hong cates not grouped samples. PCo1, PCo2 and PCo3
Kong. FEMS Microbiol Ecol 61: 496–508. explained 49% of the variability.

© 2015 Society for Applied Microbiology and John Wiley & Sons Ltd, Environmental Microbiology, 17, 3882–3897

You might also like