AEM Accepted Manuscript Posted Online 18 August 2017 Appl. Environ. Microbiol. doi:10.1128/AEM.01363-17 Copyright © 2017 American Society for Microbiology. All Rights Reserved.

1

Differences in temperature and water chemistry shape distinct diversity patterns in

2

thermophilic microbial communities

3 Cecilia M. Chiriaca,b, Edina Szekeresa,b, Knut Rudic, Andreea Baricza,b, Adriana Hegedusa,b, Nicolae

5

Dragoșa,b, Cristian Comana#

6 7

a

8

Biotechnology Department, Faculty of Biology and Geology, Babeş-Bolyai University, Cluj-

9

Napoca, Romania;

10

NIRDBS, Institute of Biological Research, Cluj-Napoca, Romania; bMolecular Biology and

c

Chemistry, Biotechnology and Food Science Department, Norwegian

University of Life Sciences, Aas, Norway.

11 12

Running title: Diversity patterns in hot spring microbial mats.

13 14

# Address correspondence to Cristian Coman, [email protected]

15 16

1

Downloaded from http://aem.asm.org/ on August 19, 2017 by FUDAN UNIVERSITY

4

17

Abstract This study presents the biodiversity and ecology of microbial mats developed in thermal

19

gradients (20-65 oC) in the surroundings of three drillings (CH, CI, MB) tapping a hyperthermal

20

aquifer from Romania. Using a metabarcoding approach, 16S rRNA genes were sequenced from

21

both DNA and RNA transcripts (cDNA) and compared. The relationships between the microbial

22

diversity and the physico-chemical factors were explored. Additionally, the cDNA data was used

23

for in silico functionality predictions, bringing new insights into the functional potential and

24

dynamics of these communities. The results showed that each hot spring determined the formation

25

of distinct microbial communities. In the CH mats (40-53 oC), the abundance of Cyanobacteria

26

decreased

27

Ectothiorhodospira, Oscillatoria and methanogenic archaea dominated the CI communities (20-

28

65 oC), while

29

Chloroflexi, Hydrogenophilus, Thermi and Aquificae. Alpha diversity was negatively correlated

30

with the increase in water temperature, while beta diversity was shaped in each hot spring by the

31

unique combination of physico-chemical parameters, regardless of the type of nucleic acid analyzed

32

(DNA vs. cDNA). The rank correlation analysis revealed a unique model that associated

33

environmental data with community composition, consisting in the combined effect of Na+, K+,

34

HCO3-, PO43- concentrations, together with temperature and electrical conductivity. These factors

35

seem to determine the grouping of samples according to location, rather than with the similarities in

36

thermal regimes, showing that other parameters beside temperature alone are significant drivers of

37

biodiversity.

38

Importance

with

the

temperature,

MB

opposite

microbial

to

that

of

(53-65 oC)

mats

Chloroflexi

were

and

mainly

Proteobacteria.

composed

of

39

Hot spring microbial mats are remarkable manifestation of life on Earth, and were

40

intensively studied for decades. Moreover, as hot spring areas are isolated and have a limited

41

exchange of organisms, nutrients and energy with the surrounding environments, hot spring

42

microbial communities can be used as model studies to elucidate the colonizing potential within

2

Downloaded from http://aem.asm.org/ on August 19, 2017 by FUDAN UNIVERSITY

18

extreme settings. Thus, they are of great importance in evolutionary biology, microbial ecology and

44

exobiology. With all the efforts that have been made, the current understanding of the influence of

45

temperature and water chemistry on the microbial community composition, diversity and abundance

46

in microbial mats is limited. In this study, the composition and diversity of microbial communities

47

developed in thermal gradients in the vicinity of three hot springs from Romania were investigated,

48

each geothermal water having particular physico-chemical characteristics. Our results expose new

49

factors that could determine the formation of these ecosystems, expanding the current knowledge in

50

this regard.

51 52

Introduction

53

Microbial mats are stratified communities of functional groups that are determined and

54

maintained by steep physico-chemical gradients (1). Hot spring microbial mats are of particular

55

interest as they are considered model ecosystems for the study of early life on Earth. By now, the

56

oldest evidence of life on land came from the ca. 3.48 billion year old Archaean rocks from Pilbara

57

Craton, Western Australia. These structures include stromatolites, or mineralized microbial mats,

58

that have developed in an ancient, hot spring environment (2). Modern hot spring habitats,

59

including various microbial mat ecosystems have been extensively studied, especially in the

60

Yellowstone National Park, USA (3 - 6), in the Tibetan Plateau, China (7 - 9), in Iceland (10, 11),

61

and Russia (12, 13). These studies have considerably improved our understanding of microbial

62

composition and diversity in terrestrial hot spring ecosystems (7). Thus, it is known that these

63

microbial mats lack seasonality and undergo changes in response to differences in physico-chemical

64

parameters, such as temperature, pH, sulfide concentrations etc. (1). These environments are

65

dominated by prokaryotes, as extreme temperatures usually exclude eukaryotic organisms (14). In

66

hyperthermal conditions that are not suitable for photosynthesis (>75oC) (15), thermal and

67

hyperthermal members of Bacteria are common inhabitants of microbial mats, being predominantly

68

affiliated to the Aquificae, Thermatogae, Deinococcus-Thermus ([Thermi]) phyla, or to certain

3

Downloaded from http://aem.asm.org/ on August 19, 2017 by FUDAN UNIVERSITY

43

thermophilic members of Proteobacteria (7). At lower temperatures and under light conditions,

70

oxygenic and anoxygenic photosynthesis usually support the primary production in the mat,

71

allowing a diversification of the microbial communities (16). Photosynthetic hot spring microbial

72

mats may be sustained by Cyanobacteria, Chloroflexi or Chlorobi, or by a combination of these

73

groups (17). Beside the bacterial domain, Archaea is also present in hot spring mats, being

74

represented especially by the Crenarchaeota and Euryarchaeota phyla (18, 19).

75

To better understand the ecology and microbial functions leading to the establishment of

76

microbial mats, a number of studies have focused on the links between the physico-chemical

77

parameters of the water and the microbial composition and diversity (7, 11, 16, 20 - 22). Within

78

those factors, temperature received much of the attention. While some studies explore the

79

temperature effect on small scale gradients (16, 20), others have selected springs within a wide

80

temperature range (11, 23). The results showed that the microbial mat composition and the

81

complexity of the ecosystem seem to be mainly determined by the rise in temperature (1). In the

82

study performed by Wang et al. (7), mat samples were collected from thermal gradients in different

83

areas and no location-specific variations were observed. Instead, the temperature was sufficient to

84

explain the variability in community composition. In contrast, other studies found that the sulphide

85

concentration (21) or the pH values (24) were important factors that control the similarity among

86

samples. Nevertheless, more effort is required to better understand the effects of physico-chemical

87

conditions on the development of hot spring microbial ecosystems.

88

As a result of geothermal resource exploitation in the second half of the 20th century, eight

89

geothermal areas were identified in Romania, with over 200 drillings ranging in depth from 800 to

90

3500 m, having the outflow water temperatures from 40oC to 120oC (25). Among these, a number

91

of drillings (80) are tapping the Pannonian hyperthermal aquifer (50 – 85oC) from the Western Plain

92

of Romania, the water that comes from 800 to 2100 m deep being used mostly for domestic heating,

93

health and recreational bathing, industrial processes or fish farming (25). Some wells have been

94

abandoned over the years and the continuous hot water flow has led to the development of specific

4

Downloaded from http://aem.asm.org/ on August 19, 2017 by FUDAN UNIVERSITY

69

mineralized communities and other photosynthetic microbial mats (26, 27). Even though the

96

biodiversity in some of these microbial mats has been previously described using 16S rDNA clone

97

libraries (26), the current study applied high-throughput sequencing techniques for a more thorough

98

investigation. It focuses on microbial mats formed along thermal gradients in hot springs from three

99

different locations, each with particular physico-chemical characteristics. An integrative approach

100

was performed that included both 16S rRNA transcripts and 16S rDNA gene sequencing. Although

101

the amount of ribosomal RNA that is detected in a microorganism is no longer viewed as a clear

102

indicator for microbial growth, a large number of copies does indicate the potential for protein

103

synthesis, for a rapid shift in metabolic function, or for an overall higher fitness of that species in

104

the ecosystem (28). The inferences of metabolic potential based on 16S rRNA transcripts may fall

105

under some degree of uncertainty (29). But, taking into consideration that extracellular RNA is less

106

stable in the environment compared with free DNA (30), functional roles can be attributed in this

107

way to intact organisms. By contrast, the DNA extracted from environmental samples is a

108

continuum of DNA derived from whole living organisms and extraorganismal, free DNA (31). A

109

more precise view on the microbial community composition, its diversity and its function may be

110

proposed by comparing the 16S rDNA gene libraries with those of 16S rRNA transcripts.

111

Thus, the aim of this research was to characterize the prokaryotic composition and

112

biodiversity in the newly investigated hot springs microbial mats associated with three hot springs

113

situated in the Western Plain of Romania and to explore their relationship with the physico-

114

chemical properties of the thermal water. As these microbial communities may bring new

115

information on the distribution of microbial groups in thermal environments, an additional effort

116

was made to compare their microbial composition with other microbial mats that have been

117

previously investigated through a metabarcoding approach.

118

Materials and Methods

119

Sampling procedure

120

Samples were taken in 2013 from microbial mats developed nearby three abandoned

5

Downloaded from http://aem.asm.org/ on August 19, 2017 by FUDAN UNIVERSITY

95

drillings from the Chiraleu (CH – drilling number 4095), Ciocaia (CI - drilling number 4045) and

122

Mihai Bravu (MB - drilling number 605) villages, at temperatures ranging from 20 to 65 oC (Table

123

1). Colorful microbial mats were observed and sampled at each location along with the temperature

124

gradient (Figure 1). In the surroundings of the Chiraleu drilling (Figure 1A), green and slimy

125

microbial mat developed at 53°C, without a clear vertical stratification (Figure 1B), as well as a

126

yellow microbial mat at 46°C which presented a typical laminated structure (Figure 1C). A thin

127

microbial mat was sampled from 40°C, with a smooth orange surface and green layers underneath

128

(Figure 1D). Three distinct microbial mats were observed in the proximity of the Ciocaia drilling

129

(Figure 1E), including green, mucilaginous and stratified microbial mats forming at 20°C (Figure

130

1F), red and slimy microbial mats developing at 35°C that do not show a clear stratification (Figure

131

1G) and white microbial mats at 65°C, more consolidated and firm near to the substrate and with a

132

mucilaginous, translucent surface layer (Figure 1H). Microbial mats and streamers were identified

133

near the Mihai Bravu drilling opening (Figure 1I), including a green, compact and smooth microbial

134

mat forming at 53°C, with thin layers (Figure 1J), a stratified microbial mat at 59°C, with an orange

135

top layer and green layers underneath (Figure 1K) and long, flexible structures forming cream -

136

colored streamers at 65°C along with the flowing hot water (Figure 1L).

137

The stable characteristics of the water sources were previously reported (32, 33) (Table 2).

138

These three drilling (CH, CI, MB) are tapping the Pannonian hyperthermal aquifer, a deposit of

139

neutral to slightly alkaline waters, with pH values ranging from 7.75 to 7.91. The high

140

concentrations of Na + (676 - 4120 mg/L), Cl- (537 - 1160 mg/L) and HCO3- (895 – 1980 mg/L) ions

141

are probably the result of the static nature of this deposit, which resulted in prolonged water - rock

142

interactions. The collecting rocks of the aquifer were formed in lagoon conditions that allowed the

143

accumulation of dissolvable chloride rocks, which ultimately impact the chemical composition of

144

the thermal deposit (32, 33). The temperature at the surface and at the bottom of each microbial mat

145

was determined using a field multiparameter (WTW Multi 340i Multiparameter, Weilheim,

146

Germany). As temperature was similar (±0.1 oC) between top and bottom (data not shown), only the

6

Downloaded from http://aem.asm.org/ on August 19, 2017 by FUDAN UNIVERSITY

121

surface values were reported. For DNA and RNA extraction, approx. 3 g of each microbial mat was

148

collected in 15 mL Falcon tubes in triplicate and TRIzolTM Reagent (Invitrogen) was immediately

149

added to fill up the tubes in order to maintain the integrity of the total RNA. All the samples were

150

transported on ice to the laboratory, and processed within the same day. After DNA and RNA

151

extraction, followed by cDNA synthesis, a total of 18 samples were used for prokaryotic cell

152

abundance estimation and metabarcoding (Table 1).

153

Nucleic acids extraction

154

Total DNA and RNA were extracted in triplicate for each microbial sample according to the

155

manufacturer’s instructions (TRIzolTM Reagent, Invitrogen) with slight modifications. 750 mg of

156

each sample (wet weight) were incubated at 60oC for 10 min, being vortexed prior to and at the end

157

of the incubation time. Then, 150 μL of chloroform were added and the tubes were vortexed for 10

158

s. The tubes were centrifuged at 10,000 × g for 10 min. The mixture was separated in three phases,

159

an upper aqueous phase, an inter-phase and a lower phenol-chloroform phase. The upper phase was

160

transferred in a new tube for RNA extraction, while the other two phases were used for DNA

161

purification. For DNA extraction, 300 μL of 100% ethanol were added to the lower phases, and the

162

components were gently mixed by inverting the tubes. Following a 3 min incubation period, the

163

tubes were centrifuged for 15 min at 12000 × g and 4°C to pellet the DNA. The pellet was

164

resuspended in1 mL of 0.1 M sodium citrate in 10% ethanol, pH 8.5, and was incubated at room

165

temperature for 30 min. The tubes were centrifuged for 15 min at 12000 × g at 4°C. This washing

166

step was repeated one more time, and the DNA pellets were resuspended in 1.5 mL of 75% ethanol.

167

The tubes were incubated for another 20 min at room temperature, and then were centrifuged for 15

168

min at 12000 × g and 4°C. The supernatant was discarded and the DNA was air dried for 15 min.

169

The pellet was resuspended in 500 μL of 8 mM NaOH by pipetting up and down. Then, the tubes

170

were centrifuged for 10 min at 12,000 × g and 4°C, the supernatant was removed, and the DNA

171

pellet was eluted in 35 µl PCR Grade water. The DNA replicates of the same sample were pooled,

172

and the DNA concentration was determined based on the Qubit® dsDNA BR Assay Kit, using a

7

Downloaded from http://aem.asm.org/ on August 19, 2017 by FUDAN UNIVERSITY

147

173

Qubit® Fluorometer (data not shown).

174

RNA extraction and cDNA synthesis For RNA purification, an equal amount of phenol – chloroform – isoamyl alcohol (25:24:1)

176

was added to the aqueous supernatant (see previous section) and the tubes were vortexed and

177

centrifuged at 10,000 x g for 2 min. The upper phase was carefully transferred to a new tube and

178

mixed with an equal amount of isopropyl alcohol. The samples were incubated at room temperature

179

for 10 min and then centrifuged at 4oC and 10,000 x g for 10 min. The supernatant was removed,

180

and 1 mL of 70% ethanol was added to the tubes placed on ice. After 5 min of incubation the tubes

181

were centrifuged at 4oC and 8,000 x g for 5 min. The ethanol was discarded and the RNA was

182

resuspended in 30 μL of PCR Grade water. Next, the DNA was removed using the TURBO DNA-

183

freeTM kit (ThermoFisher Scientific) and the RNA replicates of the same sample were pooled. The

184

RNA concentration was measured using the Qubit® Fluorometer and the Qubit® RNA BR Assay

185

Kit (data not shown). Total RNA reverse transcription was performed starting with 200 ng of

186

purified RNA for each sample using the First Strand cDNA Synthesis kit (Thermo Scientific)

187

according to the manufacturer’s instructions.

188

Quantification of prokaryotic cell abundance by qPCR

189

The PRK341F/PRK806R primers were used to target the 16S rRNA gene in the total

190

extracted DNA and cDNA by quantitative real time PCR (qPCR) (34). The reactions were

191

performed in triplicate using the SsoFast Eva Green Supermix (Bio-Rad, Hercules, CA, USA) on

192

the CFX96 Touch™ Real-Time PCR Detection System (Bio-Rad). The following components were

193

added in the reaction mix: 7 μL 1X Sso Fast EvaGreen SuperMix (Bio-Rad), 0.4 μM of the forward

194

and reverse primers, 20 ng of DNA/RNA and RNase/DNase-free water to a final volume of 14 μL.

195

The reaction program began with an initial denaturation at 98°C for 120 s, followed by 45 cycles of

196

10 s at 98°C and an annealing/elongation step of 30 s at 55°C. The gene copy number was

197

calculated by comparing the amplification results to a standard ten-fold serial dilution of known

198

quantities of recombinant plasmids (109 – 103) carrying the targeted gene. The plasmids were

8

Downloaded from http://aem.asm.org/ on August 19, 2017 by FUDAN UNIVERSITY

175

199

prepared as previously described (35). Additionally, three negative controls were analyzed in the

200

same run. A Ct cutoff value corresponding to the lowest Ct reported for the negative controls was

201

used, and the average number of gene copies detected in these controls were subtracted from the

202

sample values.

203

Preparation of sequencing libraries The V3-V4 regions of the 16S rRNA gene were amplified from the extracted DNA and

205

cDNA using the PRK341F/PRK806R primers (34) modified by addition of Illumina-specific

206

adaptors. The coverage of the primers against the SILVA reference database was 84.1% for

207

Bacteria and 58.5% for Archaea. Each PCR reaction (25 μL) contained 1X HOT FIREPol® PCR

208

mix (Solis BioDyne, Estonia), 200 nM uniquely tagged forward and reverse primers, 1 μL of

209

sample DNA and 18 μL water. The reaction conditions were: 95°C for 15 min, followed by 25

210

cycles of 95°C for 30s, 50°C for 30s, 72°C for 45s, ending with 72°C for 7 min. The amplicon

211

concentration was measured with the Qubit®dsDNA HS Assay Kit using the Qubit® Fluorometer

212

(Thermo Fisher Scientific, USA) and equal amounts of amplicons were pooled for each sample into

213

a normalized library. The pooled library concentration was measured using the PerfeCta® NGS

214

Quantification Kit for Illumina (Quanta BioSciences, USA) and the library was diluted in Tris pH

215

8.5 to a final 4 nM concentration. Sequencing was performed on a MiSeq platform (Illumina, USA)

216

using V3 sequencing chemistry with 300 bp paired-end reads.

217

Amplicon analysis

218

Raw sequence data (SRA accession number PRJNA389363) was processed and quality

219

filtered through a combination of Usearch v8 and QIIME pipelines (36, 37). Briefly, QIIME was

220

used to extract the barcodes from the sequence data, to join the forward and reverse Illumina reads

221

and to demultiplex the sequence data. Singletons removal and quality control filtering were

222

performed using the Usearch v8 pipeline, by discarding sequences with less than 350 nucleotides

223

and those with more than 0.2 total expected errors (-fastq_maxee E option of the fastq_filter

224

command). Remaining quality control filtration included de novo and reference chimera removal

9

Downloaded from http://aem.asm.org/ on August 19, 2017 by FUDAN UNIVERSITY

204

via the Usearch v8 pipeline, using the latest version of the Greengenes database ('13_8') as a

226

reference (38). The resulting Operational Taxonomic Unit (OTU) table was converted into the

227

Biological Observational Matrix (BIOM) format (39). Taxonomy was assigned using the default

228

classifier in QIIME (Greengenes) against the updated '13_8' version of the Greengenes database at a

229

97% similarity threshold. Next, the mitochondrial and plastidial sequences were filtered out of the

230

OTU table. A multiple sequence alignment was generated with PyNAST (36) in QIIME, and the

231

alignment was used to build a phylogeny with the FastTree algorithm (40).

232

Additionally, a comparison of the alpha and beta-diversity encountered in the Romanian

233

samples to other sequences from hot spring microbial mats deposited in the Sequence Read Archive

234

(SRA) was performed, including the following entries: SRA061437 (7), SRS932137, SRS932073

235

(41), PRJEB7059 (42). As in these studies different variable regions of the 16S rRNA gene were

236

amplified and sequenced, a closed-reference OTU picking was performed in QIIME (36) using a

237

97% similarity threshold. Prior to this step, sequences were first quality filtered as described in their

238

original studies. The final data set retained only those sequences that match a full-length

239

representative sequence within the last version ('13_8') of the Greengenes database.

240

Statistical analysis

241

Rarefaction was performed at a depth of 3499 sequences per sample, followed by alpha- and

242

beta-diversity estimation in QIIME (36). Euclidean distances for physico-chemical parameters were

243

computed with the vegdist function of the 'vegan' package in R, and the Mantel test (999

244

permutations) was run to evaluate the environmental impact on the microbial diversity using the

245

weighted Unifrac distances (43). Environmental physico-chemical data was log transformed prior to

246

the analysis. PCoA plots were generated for the weighted and unweighted Unifrac matrices, using

247

the cmdscale and envfit functions of the 'stats' and 'vegan' packages in R. The environmental factors

248

were fitted onto the ordination with the envfit function of the 'vegan' package in R, being scaled by

249

their correlation to the distance matrix. The significance of the sample grouping according to

250

distinct sampling sites or distinct type of investigated nucleic acids were tested by the analysis of

10

Downloaded from http://aem.asm.org/ on August 19, 2017 by FUDAN UNIVERSITY

225

251

similarity (ANOSIM) using 999 permutations (44) in QIIME. Phyla and classes with abundances

252

higher than 1% in the sequencing libraries were plotted using the inkspot function of the 'rioja'

253

package in R. The multilevel indicator OTUs (45) for each thermal spring were established using

254

the multipatt function of the 'perm' package in R.

255

Functional prediction For functionality prediction with PICRUSt (46), the OTUs were picked using an open

257

reference approach, and de-novo OTUs were removed, keeping in the final OTU-table only the

258

OTUs that had matching Greengenes IDs ('13_8'). Data in the BIOM OTU-table was normalized

259

using the 16S rRNA gene copy number for each OTU, and the normalized table was used for

260

KEGG functional orthologs predictions. In order to evaluate the PICRUSt predictions accuracy for

261

each sample, the weighted Nearest Sequenced Taxon Index (NSTI) was computed, an index that

262

describes the extent to which microorganisms from each sample are related to fully annotated

263

genomes, a value of 0.03 meaning that on average, microorganisms in the sample were predicted

264

using a relative from the same species.

265

Results and discussion

266

Abundance of prokaryotic 16S rRNA genes and transcripts

267

The abundance of prokaryotic microorganisms in the DNA and cDNA samples was

268

estimated through qPCR by targeting the 16S rRNA gene. Generally, the copy numbers in the

269

cDNA transcripts were found to be two to three orders of magnitude lower than the numbers

270

observed in the DNA sample (Figure 2). This difference may suggest that evaluation of cell

271

abundances should not be based solely on the 16S rRNA genes. Some studies proposed that up to

272

90% of the free DNA is stable in sediments and may not be affected by nucleases (47, 48). Thus,

273

the total extracted DNA most probably reflects not only the living microbes present in the

274

environment, but also the extracellular DNAs (49). Considering the instability of the free RNA, it

275

seems that it could provide a better estimation of the actual abundance of microbial cells that can

276

synthesize proteins (1, 50). In the CH samples, the number of 16S rDNA copies decreased with

11

Downloaded from http://aem.asm.org/ on August 19, 2017 by FUDAN UNIVERSITY

256

temperature from 1.1 ± 0.01 × 108 at 40oC to 3.4 ± 0.6 × 106 copies/mg at 53oC, whereas in the other

278

sampling sites no clear pattern was observed. In the CI DNA samples, the abundance of 16S rDNA

279

genes (1.2 ± 0.05 × 107 - 2.3 ± 0.27 × 108 copies/mg) were comparable with those reported in a

280

previous study focused on the microbialites described at temperatures between 32 and 65oC in the

281

proximity of the currently investigated microbial mats (27). The copy number of 16S rRNA in the

282

MB samples (2.89 ± 0.1 × 107 – 9.9 ± 0.2 × 108 copies/mg) were similar with the other two

283

locations, and no direct relationship with temperature was observed. As in the case of rDNA, the

284

16S rRNA copy number in the cDNA samples did not show a monotonic relationship with

285

temperature fluctuations in none of the investigated hot springs (Pearson rcDNA = 0.0921; rDNA = -

286

0.211; p > 0.05) (Figure 2). Whereas a linear relationship was reported in literature between the

287

abundance of specific microbial groups and the increase in temperature, this was not the case for the

288

samples within this study. Hence, more effort is needed to examine the effect of temperature on the

289

overall microbial abundance (7).

290

Microbial communities revealed by 16S rRNA gene sequencing

291

Following raw sequences quality filtration, a number of 498,539 sequences were kept in the

292

final dataset that included both the 16S rDNA and 16S cDNA libraries. Each sample was

293

represented by 3,499 to 78,280 high quality sequences that were clustered at a 97% identity

294

threshold in 639 OTUs. Figure 3 summarizes the phyla and class-level composition of samples,

295

showing that overall, up to 13.5% of the sequencing libraries were unassigned, underlining once

296

again the importance of such extreme environments in the discovery of novel microbial groups with

297

unknown features. Archaea were present in all samples except MB65r, being represented mainly by

298

aceticlastic methanogens from the Methanosaeta (31.6%) genus in CI65, together with

299

Methanoculleus (5.3%) and Methanothermobacter (3.2%) in CI35r. Regarding the bacterial

300

taxonomic composition, the groups with the highest abundances were Proteobacteria (13.2 –

301

66.2%), Cyanobacteria (up to 40.3%), Chloroflexi (up to 24%), Bacteroidetes (up to12.8%), and

302

Firmicutes (up to 11.3%). The remaining sequences were widespread affiliated across the Bacteria

12

Downloaded from http://aem.asm.org/ on August 19, 2017 by FUDAN UNIVERSITY

277

303

domain, in the Deinococcus-Thermus (or [Thermi]) (5.7 – 11.2%), Thermatogae (up to 9.03%),

304

Nitrospirae (2.6 – 15.75%), Spirochaetes (up to 5.94%), Verrucomicrobia (up to 4.32%) and

305

Aquificae (up to 10%) phyla. Chiraleu (CH) hot spring microbial mats

307

The CH samples were mainly dominated by Proteobacteria and Cyanobacteria (Figure 3).

308

The abundance of Proteobactaria increased with temperature, an inverse relationship being observed

309

for the oxygenic phototrophic group (Figure 3). While photosynthesis can take place at

310

temperatures up to 75oC (15, 51), distinct taxonomic groups are adapted to perform this function at

311

different concentrations of oxygen, at variable temperatures or light intensities by using various

312

electron acceptors and donors (52). Moderately thermophilic cyanobacteria from the genera

313

Leptolyngbya and Oscillatoria and from the order Nostocales dominated the oxygenic mats at 40

314

and 46oC and are most probably the primary producers (1). In contrast, the Cloroflexus genus

315

became dominant at 53oC (Figure 3, Supplementary Figure 1), members of this group being able to

316

assemble a fully functional photosynthetic apparatus in anoxic conditions. Such anoxic conditions

317

may be caused by the increased water temperature which lowers the solubility of various gases,

318

including oxygen (53). In the absence of O2 or in fluctuating oxic/anoxic circumstances, and in the

319

presence of H2 and H2S, Chloroflexus may grow photoautotrophically by assimilation of CO2

320

through the 3-hydroxypropionate pathway (54). In some cases, Cyanobacteria and Chloroflexi may

321

compete for limited resources (55), but their association was previously documented in thermal

322

conditions, where the oxygenic photosynthesis was dependent on sulfide consumption by

323

Chloroflexus sp. (Portillo et al., 2009; Kim, 1999). As Cyanobacteria possess chlorophyll a (680

324

nm) and Chloroflexi synthesize bacteriochlorophyll c and a (740 nm, 790 nm), they are able to

325

coexist by harvesting different wavelengths of light (52, 56). Thus, it can be assumed that the two

326

groups form distinct assemblages in the mat layers, and both can contribute to the primary

327

production of the ecosystem (52).

328

Even though the members of the Methylophilaceae family (Betaproteobacteria) are usually

13

Downloaded from http://aem.asm.org/ on August 19, 2017 by FUDAN UNIVERSITY

306

restricted to use methanol or methylamine as carbon sources and are common inhabitants of mud,

330

lake and pond water and sludge (57), this group makes almost 8% of the CH46 library. Although

331

not typical for a hot spring environment, members of this family were previously reported in wet

332

sediments collected from the Magadi Lake (Kenya), a hypersaline lake that is supplied by a series

333

of thermal springs with temperatures up to 86oC (58), and in a pond from the Kirishima geothermal

334

area, Japan (Satoh, 2013). Another abundant phylum in the CH microbial mats was Nitrospirae

335

(CH53 – 7%, CH46 – 15.75%), being mainly represented by the Nitrospira genus. It seems that the

336

moderately thermophilic conditions from CH create an environment that is suitable for nitrifying

337

organisms such as Nitrospira, a widespread group that thrives in habitats with temperatures below

338

55oC in Europe, Asia and Australia (13).

339

Ciocaia (CI) hot spring microbial mats

340

Microbial mats developed at lower temperatures (20 and 35oC) near the CI hot spring were

341

represented by almost equal proportions of Cyanobacteria (34.2 – 40.4%) and Proteobacteria (32.2

342

– 40.7%) (Figure 3). Bacteroidetes was the next most abundant group (5.6 – 6.5%) in both samples.

343

While in the sample CI20 Actinobacteria (2.8%) and Chrysiogenetes (2.9) were more abundant, in

344

the CI35 sample higher abundances of Chloroflexi (5.5%) and Firmicutes (4.2%) were detected.

345

Almost all cyanobacterial sequences in these two samples were affiliated to the Oscillatoria genus,

346

and show a decrease in abundance with the increase in water temperature (Supplementary Figure 1).

347

Usually, Oscillatoria dominated microbial mats tend to have a limited distribution, communities

348

from higher temperatures being usually formed by the Pseudoanabaena, Synechococcus and

349

Chloroflexus genera (59). Microbial mats dominated by Oscillatoria spp. were found in the waters

350

from the Chocolate Pots Hot Springs (36 to 45oC) (Yellowstone National Park, USA), rich in iron,

351

bicarbonate and silica (60). In mesothermal and saline environments, nitrogen fixing members of

352

Oscillatoria were found to be the first in colonizing new surfaces, as in the case of the microbial

353

mats from the Mellum island in the North Sea and in the Laguna Figueroa (Pacific Ocean) (61, 62).

354

Also, Oscillatoria boryana formed a predominant population in the mat developed around a sulfide-

14

Downloaded from http://aem.asm.org/ on August 19, 2017 by FUDAN UNIVERSITY

329

rich geothermal spring in Rotorua (New Zealand) (63). Within Proteobacteria, the Alpha - and

356

Gamma - classes prevailed in samples CI20 and CI35. Salinarimonas rosea and Ectothiorhodospira

357

spp. were the most dominant taxa in sample CI20, both of them being formerly encountered in hot

358

spring microbial mats. The first one is a halotolerant, facultatively anaerobic bacterium isolated

359

from a salt mine in Yunnan (China) (64), and was also encountered in a phototrophic microbial mat

360

from Hot Lake, Washington (USA) (65). The second taxon includes purple sulfur bacteria that may

361

form a red layer in microbial mats under the oxic, green layer dominated by Cyanobacteria (61, 62).

362

Like all purple sulfur bacteria, Ectothiorhodospira members are capable of photosynthesis, using

363

hydrogen sulfide or monothioarsenate instead of water as electron donors, oxidizing them to

364

elemental sulfur or arsenate, respectively (66, 67). Because the synthesis of bacteriochlorophyll in

365

these microorganisms is inhibited in the presence of oxygen and under high light intensities, they

366

are found in mats developed in the photic zone of aquatic environments at the oxic/anoxic interface

367

(68). Similar hot spring biofilms composed primarily of Oscillatoria-like and Ectothiorhodospira-

368

like organisms were previously described in Mono Lake, California (USA) (69). Other

369

microorganisms with a high abundance in the sample collected from 35oC were included in the

370

Chloronema genus (5% of the sequencing library) of the Chloroflexi phylum. This is a surprising

371

observation since these anoxygenic filamentous phototrophs usually require low temperatures (4 –

372

15oC) for growth (70). The presence of Chloronema members in CI35, that can tolerate higher

373

temperatures than those previously reported in literature, may indicate the existence of a new taxon

374

within this genus. The Flexibacter genus was also dominant (4.5%), some species (i.e. F. flexis, F.

375

roseolus, F. ruber) being common for hot spring environments (71, 72).

376

The CI65 sample comprised 101 OTUs with highly uneven proportions. The Methanosaeta

377

genus comprised 31.6% of the library. This group, even though is usually encountered in anaerobic

378

sediments and sewage sludge digesters (73), was also detected in thermal settings such as the

379

geothermal waters of the Sungai Klan spring (Malaysia) (42), and in an alkaline hot spring from

380

Iceland (74). While in the other CI samples Gamaproteobacteria were dominant, it seems that they

15

Downloaded from http://aem.asm.org/ on August 19, 2017 by FUDAN UNIVERSITY

355

381

are replaced by Betaproteobacteria at 65oC. The second abundant OTU (18.5%) was affiliated to

382

Hydrogenophilus genus of Betaproteobacteria, a group of thermophilic and aerobic organisms, that

383

can have a facultatively chemolithoautotrophic metabolism, using H2 as electron donor and CO2 as

384

a carbon source (75). Mihai Bravu (MB) hot spring microbial mats

386

The microbial mats from MB were encountered at the highest temperature range, from 53 to

387

65oC. In these microbial mats, Cyanobacteria is present as a major group only at 53oC (5.4%)

388

(Figure 3). It seems that the environmental conditions favor the Chloroflexi phylum that rises from

389

15% in MB53 and MB65, to 24% of the library at 59oC. In contrast, the proportion of

390

Betaproteobacteria (Rhodocyclales) increased with temperature (Supplementary Figure 1), being

391

mostly represented by Hydrogenophilus members. These results imply that the MB mats may be

392

sustained by either anoxigenic photosynthesis, or by chemolithoautotrophy through the oxidation of

393

hydrogen, or by both of these strategies. Compared with the CI and CH communities, some

394

particular and dominant phyla in MB samples are [Thermi] (Deinoccocus - Thermus) (5.7 – 11.2%)

395

and Aquificae (10%). Both of these phyla are a collection of bacteria adapted to harsh and extreme

396

environmental settings (76, 77). Aquificae representatives became abundant only in the MB65

397

sample, as they are strictly thermophiles with an optimum growth temperature above 65oC (78).

398

They were encountered in volcanic and geothermal heated environments, such as the hot ponds in

399

Iceland, the Yellowstone National Park, USA (79), in Tuscany, Italy (80), or in neutral and acidic

400

hot springs from Philippines (81). The vast majority of these organisms are obligate autotrophs,

401

using hydrogen and carbon dioxide as electron and carbon sources, and oxygen as final electron

402

acceptor (82). Moreover, the wide distribution of Aquificae may indicate that these bacteria have an

403

important role in high temperature habitats as primary producers (79, 82).

404

The Meiothermus and Thermus genera within the [Thermi] phylum are dominant members

405

of the MB mats. The different optimal growth temperatures for the two taxa may explain why the

406

Meiothermus proportion is higher in the MB53 and MB57 samples, whereas the Thermus prevails at

16

Downloaded from http://aem.asm.org/ on August 19, 2017 by FUDAN UNIVERSITY

385

65oC. A similar trend, a decrease of Meiothermus abundances with temperature and an increase of

408

Thermus proportion, was observed in microbial mats colonizing an aquifer bore runoff channel

409

from Australia (83). They are commonly encountered genera in hot springs with moderate to high

410

temperatures (50 - 99oC) and slightly alkaline or acidic waters, as those from Kamchatka (84),

411

Yellowstone National Park (85) or Thailand (21). Although these microorganisms do not usually

412

become major components in surface environments, situations similar to MB mats have been

413

encountered in several other springs in Iceland and Tibetan Plateau, China (7, 11).

414

Indicator species in the hot spring microbial mats

415

As previously mentioned, a number of 639 OTUs were observed in the microbial mats

416

collected from the three distinct locations. A multilevel-indicator species analysis (Supplementary

417

Table 2) revealed that only few of them (

Differences in Temperature and Water Chemistry Shape Distinct Diversity Patterns in Thermophilic Microbial Communities.

This report describes the biodiversity and ecology of microbial mats developed in thermal gradients (20 to 65°C) in the surroundings of three drilling...
2MB Sizes 1 Downloads 14 Views