Protistan diversity in a permanently stratified meromictic lake (Lake Alatsee, SW

2

Germany).1

Accepted Article

1

3 4

Andreas Oikonomou, Sabine Filker, Hans-Werner Breiner, Thorsten Stoeck*

5 6

Department of Ecology, University of Kaiserslautern, Erwin Schroedinger Str. 14, D-67663,

7

Kaiserslautern, Germany

8 9

*

Corresponding author: Thorsten Stoeck

10

address: Erwin Schrödinger Str. 14, 67663 Kaiserslautern, Germany

11

e-mail: [email protected]

12

telephone number: +49 631-2052502

13

fax number: +49 631-2052502

14 15

Running title: Eukaryotic communities in a meromictic lake

16 17 18 19 20 21 22 23 24

This article has been accepted for publication and undergone full peer review but has not been through the copyediting, typesetting, pagination and proofreading process, which may lead to differences between this version and the Version of Record. Please cite this article as doi: 10.1111/1462-2920.12666

1 This article is protected by copyright. All rights reserved.

Accepted Article

25 26 27

Summary

28 29

Protists play a crucial role for ecosystem function(ing) and oxygen is one of the strongest

30

barriers against their local dispersal. However, protistan diversity in freshwater habitats with

31

oxygen gradients received very little attention. We applied high-throughput sequencing of the

32

V9 region (18S rRNA gene) to provide a hitherto unique spatiotemporal analysis of protistan

33

diversity along the oxygen gradient of a freshwater meromictic lake (Lake Alatsee, SW

34

Germany). In the mixolimnion, the communities experienced most seasonal structural

35

changes, with Stramenopiles dominating in autumn and Dinoflagellata in summer. The

36

suboxic interface supported the highest diversity, but only 23 OTUs95% (mainly Euglenozoa,

37

after quality check and removal of OTUs with less than three sequences) were exclusively

38

associated with this habitat. Eukaryotic communities in the anoxic monimolimnion showed

39

the most stable seasonal pattern, with Chrysophyta and Bicosoecida being the dominant taxa.

40

Our data pinpoint to the ecological role of the interface as a short-term “meeting point” for

41

protists, contributing to the coupling of the mixolimnion and the monimolimnion. Our

42

analyses of divergent genetic diversity suggest a high degree of previously undescribed

43

OTUs. Future research will have to reveal if this result actually points to a high number of

44

undescribed species in such freshwater habitats.

45 46 47 48 49 50

2 This article is protected by copyright. All rights reserved.

Accepted Article

51 52 53

Introduction

54 55

Ecological processes in freshwater systems depend on species diversity and

56

contemporary functional traits of the corresponding organisms (Cardinale et al., 2011).

57

Therefore, a cornerstone for a better understanding of invidual ecosystems and their

58

function(ing) is an in-depth knowledge of the taxonomic inventory of these ecosystems. A

59

key element in freshwater ecosystems are unicellular eukaryotes (protists), which are of

60

crucial importance when it comes to energy- and carbon flow (Finlay and Esteban, 1998;

61

Sherr and Sherr, 2002). For example, small pigmented protists exert a resource-control

62

through the dependence of bacteria on photosynthetically produced carbon while mixotrophic

63

and heterotrophic protists have a predatory control and are major loss factors for bacteria

64

(Jürgens and Massana, 2008; Sanders, 2011). As prey, protists channel carbon and energy to

65

higher trophic levels. For example, they significantly contribute to the diet of a variety of

66

multicellular organisms such as Daphnia (Callieri et al., 1999). Because of the multitude of

67

distinct functional traits of individual protistan taxon groups, their differential feeding

68

preferences and responses and their distinct food-quality (Callieri et al., 1999; Glücksman et

69

al., 2010; Roberts et al., 2011; Šimek et al., 2013) the taxon composition of protistan

70

communities in individual habitats plays a crucial role for ecosystem function(ing). Even

71

though the diversity of protists in freshwater systems is presumably much higher compared to

72

ocean surface waters (Logares et al., 2009), freshwater bodies, specifically mountain lakes,

73

have received only very little attention compared to marine waters when it comes to protistan

74

plankton diversity (Triadó-Margarit and Casamayor, 2012).

75

Among freshwater lakes, stratified meromictic lakes with a pronounced oxygen

76

gradient belong to the most complex systems (oxic mixolimnion, suboxic chemocline, anoxic 3 This article is protected by copyright. All rights reserved.

sulfidic monimolimnion). Prior molecular diversity surveys have revealed the fine-scale

78

architecture of protistan plankton communities in stratified marine or brackish water bodies

79

(Stoeck et al., 2003; Stoeck et al., 2006; Stock et al., 2009; Behnke et al., 2010; Wylezich and

80

Jürgens, 2011). Because oxygen is one of the strongest barriers against local protistan

81

dispersal and gene flow (Forster et al., 2012), it is not surprising that all of these studies

82

uncovered substantial changes in protistan community structures along stratification gradients

83

with a very high diversity of previously described and undescribed protists. Only very few

84

protistan diversity studies have targeted freshwater habitats with oxygen gradients (Šlapeta et

85

al., 2005; Lepère et al., 2006; Lefèvre et al., 2007). The recurrent microscopic observations of

86

the same morphotypes in freshwater ecosystems around the world have fuelled the idea of a

87

global dispersal of protists and a relatively low global protist species richness (Finlay and

88

Esteban, 1998). The same authors even claim “there is every reason to believe that all species

89

of freshwater protozoa could eventually be discovered in one small pond”, provided this pond

90

includes the different habitats that select for distinct protists, such as oxygen gradients. If this

91

were true, detailed ecological studies in few model lakes would help to understand the

92

ecosystem function(ing) of lakes worldwide. The sequence data coming from the few

93

molecular diversity studies that targeted either oxygenated (Richards et al., 2005; Šlapeta et

94

al., 2005; Luo et al., 2011) or oxygen-depleted (Luo et al., 2005; Lefèvre et al., 2007)

95

freshwater habitats contradict these microscopy observations. However, a rejection of this

96

global dispersal / low diversity of protists hypothesis in freshwater lakes requires a recording

97

of the full deck of the protistan inventory, which also considers seasonal community

98

dynamics. Both microscopy and 18S rDNA clone-library analyses are severely limited in this

99

respect. Microscopy allows only the differentiation of taxa with morphologically distinct

Accepted Article

77

100

characters recognizable by the taxonomist; cost-intensive clone-library analyses usually detect

101

only the most abundant genes present in an environmental sample due to the relatively low

102

number of clones that can be analysed by an individual project. 4 This article is protected by copyright. All rights reserved.

Next generation sequencing (NGS) technologies, such as Illumina sequencing,

104

circumvent these shortcomings of microscopy and clone-library analyses and allow for a deep

105

sequencing of a high number of samples at relatively low costs (Shendure and Ji, 2008). Here,

106

we take advantage of this approach to provide the first molecular in-depth analyses of

107

protistan community diversity along a stratification gradient in a meromictic mountain lake in

108

three different seasons.

Accepted Article

103

109 110

Results

111 112

Study site

113

Lake Alatsee (47° 33´ 39´´ N, 10° 38´ 14´´ E) is a meromictic mountain lake with a

114

maximum depth of 35 m and an 18 ha surface area located in Allgäu, Germany. The water

115

column of the lake is characterized by an upper oxygenated mixolimnion and a permanently

116

anoxic, sulfide-rich monimolimnion below 20 meters (Fig. 1). The monimolimnion dates back

117

to ca 8 000 - 10 000 years (Fröbisch et al., 1977; Weis, 1983) and a stable chemocline at a

118

depth of 16 - 19 meters with dense populations of purple sulfur bacteria (Fritz et al., 2012)

119

separates the mixolimnion from the monimolimnion. Temperature, oxygen and hydrogen

120

sulfide concentration were measured in situ by means of an MS08 AMT probe system

121

(Analysenmeßtechnik GmbH, Germany). Profiles for lake characterization are provided in

122

Fig. 1.

123 124

V9 amplicon analyses

125

We obtained 18 644 652 raw sequences in total, 13 761 912 of which met our

126

stringency criteria as “high quality” sequences (chimeras removed). These sequences

127

clustered into 24 283 OTUs95% that could be taxonomically assigned to protists and fungi.

128

Removal of OTUs with less than three sequences resulted in 8 698 OTUs95%. Results for all 5 This article is protected by copyright. All rights reserved.

community analyses derive from data set excluding OTUs with less than three sequences. The

130

number of OTUs collapsed with decreasing sequence similarity threshold at which OTUs

131

were called (Fig. 2). The initial collapse from OTU100% to OTU95% was exponential,

132

indicating a high, possibly intraspecific, microdiversity, which is very difficult to interpret

133

taxonomically and ecologically (Stoeck et al., 2010; Dunthorn et al., 2012). From OTU95%

134

downwards, the collapse of OTU-numbers is linear with a low decrease, providing our

135

reasoning to choose OTUs called at 95% sequence similarity for statistical analyses and

136

taxonomic assignments.

Accepted Article

129

137 138 139

Diversity partitioning Simpson’s alpha diversity revealed no general pattern for specific seasons or specific

140

depths (Fig. 3). The IF harbored the lowest diversity in autumn (IF-Au). In Spring (Sp) and

141

summer (Su), the mixolimnion (MIXO-Sp and MIXO-Su) showed similar diversity values to

142

the monimolimnion (MONO-Sp and MONO-Su, respectively).

143

The stratification gradient in the lake’s water column rather than seasonal patterns is

144

mirrored in the Jaccard (Jabundance)-UPGMA distance, which revealed three discrete clusters

145

(Fig. 4). Identical clusters were also recovered with the incidence based Jaccard index (J binary)

146

(Fig. S1). All communities from the oxic mixolimnion (MIXO-Au, MIXO-Sp and MIXO-Su)

147

were united in one cluster and were clearly separated from the cluster of the anoxic samples

148

(MONO-Au, MONO-Sp and MONO-Su). The MIXO-Sp and MIXO-Su protistan

149

communities are more similar to each other than to the MIXO-Au community, confirming the

150

Simpson’s alpha-diversity pattern of these three communities (see above, Fig. 3). The only

151

exception disrupting the stratification gradient pattern is the IF protistan community collected

152

in spring (IF-Sp), which appeared relatively similar to the MONO-Sp community, and

153

clusters within the anoxic MONO-clade. The reasoning for this clustering can be found in the

154

following numbers: 71% of all OTUs detected in IF-Sp were also found in the MONO-Sp 6 This article is protected by copyright. All rights reserved.

protistan community. In comparison, IF-Sp and IF-Su share 51%, whereas IF-Sp and IF-Au

156

share only 42%.

Accepted Article

155

157 158 159

Taxonomic protistan community composition In terms of major taxonomic groups, there is only moderate variation among depths

160

and among seasons (Fig. 5). An interesting observation on higher taxon level is the

161

dominance of Ciliophora, Dinoflagellata, Stramenopiles and Cryptophyta in the mixolimnion

162

(MIXO) and interface (IF) at all seasons with other taxon groups (e.g. Amoebozoa,

163

Euglenozoa, Fungi, Chlorophyta, Metamonada) being less diverse. Alveolata OTUs

164

(Ciliophora and Dinoflagellata) peak in the summer samples in the MIXO (68% of all OTUs

165

in this sample) and in the IF (51%) with specifically Dinoflagellata showing a notable

166

increase in the summer samples. In the anoxic sulfidic monimolimnion (MONO), the relative

167

proportion of Stramenopiles OTUs is still in the same order of magnitude as in the other two

168

sample depths, but the relative proportion of OTUs in other taxon groups is more evenly

169

distributed. For example, Alveolata account for only 30% maximum (summer sample).

170

Because spatial structuring along the stratification gradient is a stronger force shaping

171

community patterns rather than seasonal changes (see above, Fig. 4) we analyzed the protistan

172

communities on higher taxonomic resolution by depths pooled over the three seasons (Fig. 6).

173

Differences in community structures in the MIXO, IF and MONO became evident, when

174

identifying and visualizing the taxonomic OTU distribution that is unique to each of these

175

depths. We identified 412 OTUs exclusively present in the MIXO (=15% of all OTUs

176

observed only in the MIXO), only 23 OTUs unique in the IF (=6% of all OTUs observed only

177

in the IF) and 111 OTUs in the anoxic sulfidic MONO (=22% of all OTUs observed only in

178

the MONO). Ciliophora accounted for 52% of the exclusive OTUs in the MIXO, followed by

179

Stramenopiles (22%) and Dinoflagellata (18%) (Fig. 6A). A similar picture emerged for the

180

IF regarding Ciliophora (52%) and Stramenopiles (26%) (Fig. 6B). However, in the MIXO, 7 This article is protected by copyright. All rights reserved.

the vast majority of exclusive ciliates are Spirotrichea (Fig. 6A). This is different in the IF, in

182

which the proportion of exclusive colpodean and protostomatean OTUs is as large as the

183

proportion of spirotrichs (Fig. 6B). Furthermore, in contrast to the MIXO, no unique

184

Chlorophyta and Centrohelida were found. Therefore unique Euglenozoa (2 Kinetoplastea

185

OTUs and 1 Euglinida OTU) occurred in the IF together with 1 unique OTU assigned to

186

Peridiniophycidae (Dinoflagellata) and 1 Cryptophyta OTU. The relative taxonomic

187

distribution of the exclusive OTUs in the anoxic sulfidic MONO decisively differed from both

188

above water compartments (Fig. 6C). Exclusive Amoebozoa accounted for 27%, followed by

189

Metamonada (21%). Exclusive Chlorophyta accounted for 11% in the MONO. Exclusive

190

OTUs assigned to Cercozoa (6%), Fungi (4%), and Apusomonadida (2%) occurred only in

191

MONO. The proportion of exclusive Ciliophora, Dinoflagellata and Stramenopiles OTUs

192

decreased dramatically in the MONO compared to the MIXO and IF.

Accepted Article

181

193 194

Divergent genetic diversity

195

The 200 most abundant OTUs95% pooled over depths and seasons (Fig. 7) were

196

dominated by Ciliophora OTUs (n=56) and Stramenopiles (n=55), followed by Dinoflagellata

197

(n=23), Chlorophyta and Fungi (each n=12), Cryptophyta (n=11) and Cercozoa (n=7).

198

Numerous other taxon groups were included with less than five OTUs. Basically, the diversity

199

detected in these 200 most abundant OTUs95% can be categorized as follows:

200

a) OTUs that are 95%-100% similar to deposited sequences of described taxa (blue-

201

dark green). In total, 80 OTUs (40%) in our analysis fell into this category. Nearly half of the

202

ciliate OTUs (n=28), nearly half of all Dinophyceae OTUs (n=11) and more than half of all

203

Chlorophyta OTUs (n=7) in this analysis. Of the 55 Stramenopiles, 17 are 95-100% similar to

204

deposited sequences of described taxa, almost all of which are Chrysophytes. But only 4

205

Cryptophyta, 3 Fungi and 1 Cercozoa OTU fell into this category.

8 This article is protected by copyright. All rights reserved.

b) OTUs that are 85%-94.9% similar to deposited sequences of described taxa

Accepted Article

206 207

(yellow-green). In total 104 of the 200 most abundant OTUs fell into this biggest category.

208

c) OTUs that are 80%-84.9% similar to deposited sequences of described taxa (red-

209

orange). In this category of highly divergent sequences we found 16 out of 200 OTUs

210

consisting of five Stramenopiles, 2 Dinoflagellata, 2 Cryptophyta, 2 Cercozoa and one of each

211

Apicomplexa, Kinetoplastida, Fungi, Heterolobosea and Jakobida. Interestingly, none of the

212

numerous abundant Ciliophora and none of the abundant Chlorophyta OTUs fell into this

213

category.

214 215

Discussion

216 217

Seasonal patterns of protistan plankton and their spatial distribution along stratification

218

gradients are well documented from molecular as well as from microscopy studies (Fenchel et

219

al., 1990; Behnke et al., 2006; Saccà et al., 2008; Stock et al., 2009; Behnke et al., 2010;

220

Wylezich and Jürgens, 2011; Esteban et al., 2012; Stoeck et al., 2014). However, the depth of

221

(sequence) sampling conducted in a permanently stratified freshwater meromictic lake in a

222

temporal and spatial resolution is hitherto unique. Therefore, the data obtained in this study

223

provided new insights into an only little-studied ecosystem.

224

The protistan community in the oxic mixolimnion experienced most structural changes

225

during seasons. This is not unexpected because this water layer is mostly affected by

226

atmospheric changes such as temperature and light availability, but also by seasonal

227

succession of zooplankton affecting protistan community structures (Miracle et al., 1992;

228

Klaveness and Løvhøiden, 2007; Stewart et al., 2009). The dominant role of stramenopiles

229

has been identified in oligotrophic mixed freshwater bodies (Richards et al., 2005), including

230

the mixolimnion of meromictic lakes (Tarbe et al., 2011). The ecophysiological versatility of

231

Dinoflagellata makes them a very successful and abundant component in freshwater 9 This article is protected by copyright. All rights reserved.

ecosystems (Smayda, 2002; Taylor et al., 2008), which also applies to meromictic lakes. The

233

strong seasonal fluctuations of Dinoflagellata with summer-peaks may seem surprising at first

234

sight, because Dinoflagellata can adapt to low-light conditions (autumn and winter) and

235

maintain maximum cell division rates (Chan, 1978; Jakobsen et al., 2000). However, under

236

nutrient-limitation as in the oligotrophic mixolimnion of lake Alatsee, low-light adaptation in

237

Dinoflagellata may fail (Prézelin and Matlick, 1983). Furthermore, increasing temperatures

238

are likely to support dinoflagellate diversity peaks in summer. Even though freshwater

239

dinoflagellates adapted to cold-water are well known (e.g. Borghiella dodgei with optimal

240

growth rate at 5°C, (Flaim et al., 2012)), a large proportion of Dinoflagellata in lake Alatsee

241

seems to appear only in higher abundances at elevated temperatures. One example is

242

Peridinium, which was also detected in the mixolimnion of lake Alatsee, showing an optimal

243

growth at 23°C (Lindström, 1984). Representatives of the class Spirotrichea dominated the

244

exclusive Ciliophora found in the lake’s mixolimnion. Ciliophora is one of the most diverse

245

eukaryotic taxon groups occurring in a variety of aquatic habitats (Foissner et al., 2008).

246

Within this phylum, the class Spirotrichea (mainly oligotrichids) constitute a quantitatively

247

important fraction of the ciliate population in the epilimnion of freshwater lakes persisting

248

throughout all seasons (James et al., 1995; Zingel and Ott, 2000; Pfister et al., 2002).

249

Spirotrichea are gradually replaced from the hypolimnion of freshwater lakes when oxygen

250

levels drop (James et al., 1995; Zingel, 2005), such as in the suboxic interface.

Accepted Article

232

251

Seasonal variation is less pronounced in the suboxic interface, where Ciliophora

252

dominated the planktonic protists in autumn and summer. Through a variety of lifestyles,

253

ciliates belong to the most successful protistan taxon groups that can cope with low-oxygen

254

conditions or anoxia (Fenchel and Finlay, 1995 and references within). A motile response of

255

many ciliates to oxygen sensing results in an aggregation of these organisms at oxic-anoxic

256

boundary layers in aquatic habitats where they find their preferred oxygen tension (Fenchel

257

and Finlay, 2008). This finding is congruent with the dominance of ciliates at natural oxic10 This article is protected by copyright. All rights reserved.

anoxic boundary layers in molecular studies (Šlapeta et al., 2005; Behnke et al., 2006; Stock

259

et al., 2009). It was an interesting observation that the interface supports the highest overall

260

protistan diversity, but at the same time only very few OTUs were exclusive to this habitat.

261

Furthermore, in terms of taxon composition, the interface displays the highest seasonal

262

dynamic (Fig. 4). These data point to the special ecological importance of boundary layers in

263

meromictic lakes. The interface is inhabited by dense populations of purple sulfur bacteria

264

(PBS) including Chromatium, Lamprocystis, Thiocystis and Thiodyction (Fritz et al., 2012),

265

all of which are typical for such chemoclines in the photic zone of stratified lakes (Storelli et

266

al., 2013). Investigations in a comparable chemocline at Lake Cadagno in Switzerland

267

showed that these bacteria are the ecosystem’s key players in inorganic carbon

268

photoassimilation (Camacho et al., 2001). Likewise, other products of anaerobic bacterial

269

metabolism originating from the sediment and the monimolimnion such as sulfide, methane

270

and ammonia, also support a rich and abundant bacterial community in the chemocline of

271

meromictic lakes (Cloern et al., 1983; Lüthy et al., 2000; Pimenov et al., 2003). The variety

272

of chemoautotrophs in the chemocline then resupplies several oxidants such as Fe3+, NO2- and

273

NO3- and S0 to the anoxic monimolimnion (for a model of the processes see Esteban et al.,

274

2012). Even though no comparable studies exist for Lake Alatsee, it is reasonable to assume

275

that the same applies to the bacteria and microbial processes in this lake’s chemocline. Thus,

276

despite the small relative volume of the chemocline in meromictic lakes, this layer is the

277

motor for fluxes, the lynchpin for metabolic products and connects the different water layers

278

with each other. The role of protists in these processes, however, is unknown. Our data,

279

showing that the interface in lake Alatsee is a short-term “meeting point” for protists residing

280

in the mixolimnion and protists inhabiting the anoxic monimolimnion, suggest that protists

281

may play a pivotal role in these processes and also in the coupling of the different water

282

layers. Experiments with e.g. stable isotopes could confirm the role of protists in these

283

ecosystem processes.

Accepted Article

258

11 This article is protected by copyright. All rights reserved.

Only one major eukaryote evolutionary lineage occurred exclusively in lake Alatsee’s

285

chemocline, namely Euglenozoa (Fig. 6). This corroborates with previous findings: Tuomi et

286

al. (1997) identified dense Euglena populations immediately on top of a chemocline in a

287

meromictic Norwegian lake as responsible for a green water color. Such phototrophic

288

euglenids are frequent inhabitants of chemoclines under reduced light levels and can tolerate

289

sulfide concentrations (Klaveness and Løvhøiden, 2007). Orsi and colleagues (2011) provided

290

evidence on the apparent habitat specialization of Euglenozoa in suboxic habitats. Likewise,

291

Cryptophyta often form considerable populations at the oxic-anoxic transition zone of

292

stratified water masses (Gasol et al., 1993). In contrast to euglenids, cryptophytes migrate

293

further into the anoxic and sulfidic monimolimnion in order to reduce predation losses

294

(Pedrós-Alió et al., 1995). Their occurrence in the upper mixolimnion is less frequent and is

295

primarily controlled by predation (Gervais, 1998), explaining the exclusive observation of

296

Cryptophyta in the chemocline and the monimolimnion (Fig. 6).

Accepted Article

284

297

The eukaryotic communities in the anoxic monimolimnion showed the most stable

298

seasonal pattern (Fig. 5), corroborating well with the stable hydro-physicochemical conditions

299

in this deeper water layer throughout the different seasons. In addition, the lack of seasonally

300

fluctuating multicellular grazers in such permanently anoxic and sulfidic water bodies most

301

likely contributes to maintaining a seasonally relatively stable community with little

302

dynamics, successions and fluctuations. Stramenopile diversity, mainly Chrysophyta and

303

Bicosoecida, dominated the protistan communities in the anoxic monimolimnion of lake

304

Alatsee. This was unexpected because previous microscopy and molecular diversity analyses

305

found ciliates as main components in anoxic compartments of stratified water columns

306

(Finlay et al., 1996; Guhl et al., 1996; Behnke et al., 2010; Charvet et al., 2012), even though

307

Chrysophyta and Bicosoecida are frequent and abundant members in oxygen-depleted

308

habitats (Luo et al., 2005; Behnke et al., 2006; Stock et al., 2009; Wylezich and Jürgens,

309

2011). The ecological importance of phagotrophic stramenopiles as a trophic link and their 12 This article is protected by copyright. All rights reserved.

potential involvement in regulating bacterial populations in anoxic marine systems has been

311

discussed earlier (Stoeck et al., 2007; Orsi et al., 2011; Massana et al., 2013). Even though

312

morphologically and phylogenetically marine stramenopiles are clearly distinct from

313

freshwater stramenopiles (Park and Simpson, 2010), their ecological importance in anoxic

314

freshwater habitats is very likely the same.

Accepted Article

310

315

The lower diversity of Ciliophora in lake Alatsee’s anoxic monimolimnion is most

316

likely associated with their low grazing potential and efficiency (Oikonomou et al., 2014). It

317

seems that the prevailing bacterial community structure in lake Alatsee’s monimolimnion

318

does not allow the growth of diverse and large ciliate populations (despite the numerous

319

adaptation mechanisms of ciliates to anoxia, Fenchel and Finlay, 1995), compared to the

320

much smaller bicosoecid and chrysophyte flagellates. Thus, biotic interactions also seem to

321

play an important role shaping protistan plankton community structures in the anoxic water

322

bodies of meromictic lakes. The classes Armophorea and Karyorelictea of Ciliophora were

323

exclusively found interannually in the anoxic samples. These groups typically have a lifestyle

324

adapted to anoxic conditions (Fenchel and Finlay, 1995; Lynn and Small, 2002), explaining

325

their exclusive residency in anoxic habitats of stratified water columns (Stock et al., 2009).

326

Additionally, the exclusive occurrence of Amoebozoa (mainly archamoebae) and

327

Metamonada (mainly the free-living diplomonads) in the anoxic monimolimnion is rooted in

328

these organism lifestyles and anaerobic metabolism (Bringaud et al., 2010). These occur very

329

rarely in oxygenated freshwater ecosystems and are frequent inhabitants in oxygen-depleted

330

freshwater (Bernard et al., 2000). The exclusive OTUs of Fungi in the anoxic compartment

331

was less surprising. They are well documented from anoxic habitats (Luo et al., 2005; Orsi et

332

al., 2011), and recent experiments show evidence for fungal anaerobic metabolism (such as

333

denitrification, Stief et al., 2014).

334

The appearance of Chlorophyta in the lake’s dark zone is not unusual. Some members

335

of the Chlorophyta (e.g. Chlamydomonas sp.) are capable of producing oxygen in the light, 13 This article is protected by copyright. All rights reserved.

but turn rapidly to anaerobic metabolism after exposure to anoxia and darkness (Meuser et al.,

337

2009; Atteia et al., 2013). Moreover, several green algae are able to tolerate anoxic conditions

338

in the monimolimnia of meromictic lakes for an extended period of time (Klaveness and

339

Løvhøiden, 2007), while others can switch to a phagotrophic lifestyle (Bell and Laybourn-

340

Parry, 2003; Maruyama and Kim, 2013). In the anoxic monimolimnion Chlorophyta may find

341

a refuge from the predation they experience in the photic zones of lakes (Arvola et al., 1992).

Accepted Article

336

342

Our results indicate that freshwater anaerobic communities are different from the

343

anaerobic communities in brackish environments (Stock et al., 2009; Behnke et al., 2010), in

344

saline meromictic lakes (Charvet et al., 2012) and in stratified marine water columns (Stoeck

345

et al., 2003; Stoeck et al., 2006). Our network analyses (Fig. 7) revealed a high degree of

346

genetic divergence when comparing our obtained V9 18S rDNA amplicons to deposited gene

347

data. Such observations are not unusual, when it comes to low-abundant taxa (“rare OTUs”),

348

because this category of organisms usually refers to the microbes that present themselves only

349

very rarely under the microscope (Lynch et al., 2012). However, we find such a high genetic

350

divergence even in the most abundant OTUs. Most of this genetic divergence lies in the

351

cercozoans and the stramenopiles. But also in the ciliates, a taxon group that belongs to the

352

morphologically best-characterized protists because of intense microscopy research for almost

353

two centuries (Ehrenberg, 1838), we still find substantial genetic divergence.

354

It remains unclear whether we are dealing with hitherto novel eukaryotic lineages or

355

described but unsequenced eukaryotic groups. Taxonomic boundaries differ among various

356

eukaryotic groups and a unified classification for 18S genera does not exist (Caron et al.,

357

2009). Phylogenetic analyses of full length sequences are needed to accurately describe such

358

hidden diversity and would help to uncover the unsequenced or novel status of such

359

phylotypes. One approach would be the design of specific reverse primers targeting selected

360

SWARMs, which in combination with more universal (but still group-specific) forward

361

primers would produce the near complete 18S rDNA fragment for a solid phylogenetic 14 This article is protected by copyright. All rights reserved.

analysis. Based on this fragment, specific FISH-probes could be designed for the hunt of the

363

corresponding morphotype in fixed sample material (Kolodziej and Stoeck, 2007; Massana

364

and Pedrós-Alió, 2008). Previous hunts for organisms behind novel genes nicely demonstrate

365

that in bacteria (Lynch et al., 2012) as well as in protists novel genes often correspond to

366

novel organisms (Kolodziej and Stoeck, 2007; Massana and Pedrós-Alió, 2008).

Accepted Article

362

367

Several approaches have been suggested in the literature of how to access the

368

organisms behind novel genes and to reveal their taxonomic identity. Such approaches include

369

the design of specific oligonucleotide probes targeting the new gene-bearing organisms

370

microscopically (Kolodziej and Stoeck, 2007); or the single-cell sequencing of lugol-fixed

371

and photo-documented cells from environmental samples (Auinger et al., 2008; Stoeck et al.,

372

2014). Comparing genes with morphology is under the current circumstances more than

373

difficult (for a detailed discussion we refer to Stoeck et al., 2014). A promising approach

374

towards this goal is the comparison of taxonomic marker genes like the hypervariable V4

375

region of the SSU rDNA (Pawlowski and Holzmann, 2014) in combination with deep-

376

sequencing of as many freshwater habitats as possible. The tools are at hand and affordable.

377

Even though taxonomic assignments of most of the obtained OTUs at a species-level are still

378

not possible, novel and powerful computational tools are available for a multiple-sample

379

comparison (Bittner et al., 2010; Caporaso et al., 2010; Barberán et al., 2012; Lynch et al.,

380

2012) allowing the inference of geographical patterns of OTUs.

381 382

Experimental procedures

383 384 385

Sample collection Water samples were collected with a 5 l Niskin bottle (Hydro-Bios GmbH, Germany)

386

from the mixolimnion (6-7 meters), the interface (18-19 meters) and the anoxic

387

monimolimnion (22-23 meters) of lake Alatsee in autumn 2011 (October 27), spring 2012 15 This article is protected by copyright. All rights reserved.

(May 29) and summer 2012 (August 28). To avoid contact with the atmosphere, samples from

389

the suboxic interface and the anoxic monimolimnion were directly transferred to evacuated

390

and sterile 3 l Ethyl Vinyl Acetate bags (EVA bags, Baxter UK). Cells from each sample were

391

collected on 0.65 µm Durapore membranes (Millipore Co.) under gentle pressure (< 50 ml

392

min-1) using a peristaltic pump (Ecoline ISM 1079, Ismatec, Germany). The filtration step

393

was carried out in the field immediately after sampling. Depending on the cell concentration

394

in each sample, 1.2 - 2.5 l of water was filtered per sample. To avoid oxygen contamination,

395

the interface and monimolimnion samples were filtered by connecting the filtration system

396

directly to the EVA bags. Three replicate membrane filters were obtained for each layer and

397

each season. The membranes were stored in RNAlater (Qiagen GmbH, Germany) for 24 h at

398

room temperature and then were frozen at - 80°C until further processing in the lab.

Accepted Article

388

399 400

DNA extraction, PCR amplification and Illumina sequencing

401

Each filter was cut with a sterile scalpel into smaller pieces and transferred to a Lysis

402

Matrix E tube (MP Biomedicals, Germany), followed by the addition of 600 µl RLT-Buffer

403

and 6 µl β-mercaptoethanol. Filters were then shaken at 30 Hz for 45 s using a mixer mill

404

(MM200, Retsch, Germany). The tubes were centrifuged (14 000 rpm, 3 min), the supernatant

405

was retained and DNA extraction followed the protocol of Qiagen GmbH (Hilden, Germany)

406

All Prep DNA/RNA Mini kit (after step 4).

407

The concentration of bulk DNA was measured spectrophotometrically (NanoDrop

408

2000, Thermo Scientific, Wilmington, DE, USA). The hyper-variable V9 region of the 18S

409

rRNA gene was amplified using the primers 1391F (5´-GTACACACCGCCCGTC-3´ (Lane,

410

1991), S. cerevisiae position 1629-1644) and EukB (5´-GATCCTTCTGCAGGTTCACCTAC

411

-3´ (Medlin et al., 1988), S. cerevisiae position 1773-1797). The highly conserved nature of

412

the 1391F primer appears to recover more taxa compared to eukaryotic-specific primers

413

targeting the same gene region (Amaral-Zettler et al., 2009). Primers were synthesized by 16 This article is protected by copyright. All rights reserved.

Biomers (Ulm, Germany) with a hexamer sample-specific identifier tagged to the primer

415

sequence (Table S1).

Accepted Article

414

416

The PCR mixture, containing 1 U of Phusion High Fidelity DNA Polymerase (New

417

England Biolabs), was heated to 98°C for 30 s and the V9 region was amplified by 27 cycles

418

consisting of 98°C for 10 s, 60°C for 30 s and 72°C for 30 s, followed by a final 10-min

419

elongation step at 72°C. To minimize potential PCR amplification bias, we performed 3

420

independent 50 µL PCR reactions for each layer and each season. The PCR products from all

421

amplifications were visualized by agarose gel electrophoresis (2%) under UV light. The three

422

PCR reactions from the same layer of the same season were pooled together and purified

423

using the MiniElute® Reaction Cleanup kit (Qiagen GmbH, Germany).

424

All V9 tags were sequenced with the Illumina MiSeq platform from the forward (5´-

425

end) and the reverse (3´- end) primer and the paired-end reads generated from the same

426

amplicon were merged together with FLASH-1.2.4 (Magoč and Salzberg, 2011) by LGC

427

Genomics GmbH (Germany). The sequence tags have been deposited to NCBI’s Sequence

428

Read Archive (SRA) and can be found under the accession number SRP044320.

429 430 431

Sequence data processing Sequences with inaccurate hexamer identifiers, with one or with both primers incorrect

432

or incomplete and those containing one or more ambiguous nucleotides (Ns) were discarded

433

as “low quality” sequences, using QIIME v.1.7.0 (Caporaso et al., 2010). All sequences

434

starting with reverse primers were reverse complemented using a custom script. Sequences

435

with a length of 80 - 200 nucleotides, after trimming the primers were grouped into

436

Operational Taxonomic Units (OTUs) called at 100% to 89% sequence similarity in 1%-steps

437

using a modified pipeline script of usearch quality filter (usearch_qf, Edgar et al., 2011). In

438

detail, the modified script (available upon request from the authors) dereplicated the

439

sequences, sorted them by decreasing abundance and performed a de novo chimera check 17 This article is protected by copyright. All rights reserved.

using the UCHIME algorithm (Edgar et al., 2011). The detected chimeric sequences were

441

discarded and the remaining non-chimeric sequences were clustered using Uclust (Edgar et

442

al., 2010). The longest sequence of each cluster was picked as representative sequence of each

443

OTU for the following taxonomic analysis.

Accepted Article

440

444

Taxonomic affiliation for OTUs called at 95% sequence similarity (OTU95%, reasoning

445

see above) was performed with the software tool JAguc (Nebel et al., 2011) as described in

446

(Stoeck et al., 2010) using the non-redundant nucleotide NCBI database release 197.0. Non-

447

target OTUs assigned to Archaea, Bacteria, Metazoa or Embryophyta were discarded from

448

downstream analyses. The classification of major protistan taxon groups followed Adl et al.,

449

(2012).

450 451 452

Community analyses Indices of alpha diversity and similarities among the samples (beta diversity) were

453

calculated with QIIME v.1.7.0. We used the Simpson’s index of diversity (1-D), as it provides

454

robust quantification and meaningful comparison of the microbial diversity in molecular data

455

sets (Haegeman et al., 2013). The Jaccard index was used as a measure of similarity between

456

the samples based on both abundance (Jabundance) and incidence (Jbinary). Prior to beta diversity

457

estimation, the number of sequences per sample was rarefied to the smallest sample by using

458

the single_rarefaction.py-script in QIIME. Jaccard similarity values were transformed to

459

distance matrixes for an Unweighted Pair Group Method with Arithmetic Mean (UPGMA)

460

cluster analyses. Bootstrap values were calculated to measure the robustness of the UPGMA

461

dendrogramms using the jackknifed_beta_diversity.py-script and considering a total of 1 000

462

support trees. Because community statistics predominantly relies on abundant taxa or OTUs

463

(see review Dunthorn et al., 2014), we here conducted community analyses with data sets

464

excluding OTUs that include less than three sequences to accelerate computation time for

465

millions of sequence reads. Previous molecular diversity studies with amplicon data sets and 18 This article is protected by copyright. All rights reserved.

statistical analyses have already convincingly and in detail shown that diversity patterns at the

467

community scale are not affected by the removal of rare OTUs (Gobet et al., 2010; Pommier

468

et al., 2010; Zinger et al., 2012).

Accepted Article

466

469 470 471

Analyses for the detection of divergent genetic diversity To identify divergent genetic protistan diversity in lake Alatsee, the cleaned Illumina-

472

V9-reads of all depths and layers were combined, dereplicated and grouped using the

473

SWARM clustering method (Mahé et al., 2014, https://github.com/torognes/swarm).

474

SWARM group reads were based on a selected maximum number of nucleotide differences

475

(in this study d=1). Seed sequences (i.e. most abundant sequence of each swarm) were culled

476

from the biggest 200 target swarms and subjected to BLAST analyses against Genbank´s nr

477

nucleotide database (v. 201.0) to infer their taxonomic identity. The seed sequences of each

478

swarm were then aligned with Seaview (Galtier et al., 1996), prior to calculating sequence

479

similarities between each of the seed sequences using the custom script PairAligner (provided

480

by Dr. Markus Nebel, University of Kaiserslautern). The igraph R package (Csárdi and

481

Nepusz, 2006, R Development Core Team 2008) was used to build the network based on the

482

sequence similarity values. In this network two nodes were connected by an edge if they

483

shared a sequence similarity of at least 90%. The resulting network was visualized and

484

modified with Gephi v.0.8.2-beta (Bastian et al., 2009) according to the swarms’ taxonomic

485

affiliation and BLAST hit value.

486 487

Acknowledgments

488 489

This research was supported by grant STO414/10-1 of the Deutsche

490

Forschungsgemeinschaft (DFG) to TS. We are grateful to Maria Pachiadaki and Maria

491

Siegesmund for their valuable help in field sampling, to Lucie Bittner and Josef Schüle for 19 This article is protected by copyright. All rights reserved.

modifying the script usearch_qf in order to process large Illumina datasets. Katharina Zweig

493

is acknowledged for helping with estimations of the overlapping percentages of OTUs among

494

samples. We would also like to thank the authorities of the city of Füssen, Germany, for

495

permission to sample lake Alatsee.

Accepted Article

492

496 497

References

498

Adl, S.M., Simpson, A.G., Lane, C.E., et al. (2012) The revised classification of eukaryotes.

499

J Eukaryot Microbiol 59: 429-493.

500

Amaral-Zettler, L.A., McCliment, E.A., Ducklow, H.W. and Huse, S.M. (2009) A method for

501

studying protistan diversity using massively parallel sequencing of V9 hypervariable regions

502

of small-subunit ribosomal RNA genes. PLoS ONE 4: e6372.

503

Arvola, L., Salonen, K., Kankaala, P. and Lehtovaara, A. (1992) Vertical distributions of

504

bacteria and algae in a steeply stratified humic lake under high grazing pressure from

505

Daphnia longispina. Hydrobiologia 229: 253-269.

506

Atteia, A., van Lis, R., Tielens, A.G.M. and Martin, W.F. (2013) Anaerobic energy

507

metabolism in unicellular photosynthetic eukaryotes. Biochim Biophys Acta 1827: 210-223.

508

Auinger, B.M., Pfandl, K. and Boenigk, J. (2008) Improved methodology for identification of

509

protists and microalgae from plankton samples preserved in Lugol's iodine solution:

510

combining microscopic analysis with single-cell PCR. Appl Environ Microbiol 74: 2505-

511

2510.

512

Barberán, A., Bates, S.T., Casamayor, E.O. and Fierer, N. (2012) Using network analysis to

513

explore co-occurrence patterns in soil microbial communities. ISME J 6: 343-351.

514

Bastian, M., Heymann, S. and Jacomy, M. (2009) Gephi: An open source software for

515

exploring and manipulating networks. International AAAI Conference on Weblogs and Social

516

Media.

20 This article is protected by copyright. All rights reserved.

Behnke, A., Barger, K.J., Bunge, J. and Stoeck, T. (2010) Spatio-temporal variations in

518

protistan communities along an O2/H2S gradient in the anoxic Framvaren Fjord (Norway).

519

FEMS Microbiol Ecol 72: 89-102.

520

Behnke, A., Bunge, J., Barger, K., Breiner, H.W., Alla, V. and Stoeck, T. (2006)

521

Microeukaryote community patterns along an O2/H2S gradient in a supersulfidic anoxic fjord

522

(Framvaren, Norway). Appl Environ Microbiol 72: 3626-3636.

523

Bell, E.M. and Laybourn-Parry, J. (2003) Mixotrophy in the antarctic phytoflagellate,

524

Pyramimonas gelidicola (Chlorophyta: Prasinophyceae). J Phycol 39: 644-649.

525

Bernard, C., Simpson, A.G.B. and Patterson, D.J. (2000) Some free-living flagellates

526

(protista) from anoxic habitats. Ophelia 52: 113-142.

527

Bittner, L., Halary, S., Payri, C., Cruaud, C., de Reviers, B., Lopez, P. and Bapteste, E.

528

(2010) Some considerations for analyzing biodiversity using integrative metagenomics and

529

gene networks. Biol Direct 5: 47.

530

Bringaud, F., Ebikeme, C. and Boshart, M. (2010) Acetate and succinate production in

531

amoebae, helminths, diplomonads, trichomonads and trypanosomatids: common and diverse

532

metabolic strategies used by parasitic lower eukaryotes. Parasitology 137: 1315-1331.

533

Callieri, C., Pugnetti, A. and Manca, M. (1999) Carbon partitioning in the food web of a high

534

mountain lake: from bacteria to zooplankton. J Limnol 58: 144-151.

535

Camacho, A., Erez, J., Chicote, A., Florín, M., Squires, M.M., Lehmann, C. and Backofen, R.

536

(2001) Microbial microstratification, inorganic carbon photoassimilation and dark carbon

537

fixation at the chemocline of the meromictic Lake Cadagno (Switzerland) and its relevance to

538

the food web. Aquat Sci 63: 91-106.

539

Caporaso, J.G., Kuczynski, J., Stombaugh, J., et al. (2010) QIIME allows analysis of high-

540

throughput community sequencing data. Nat Methods 7: 335-336.

541

Cardinale, B.J., Matulich, K.L., Hooper, D.U., et al. (2011) The functional role of producer

542

diversity in ecosystems. Am J Bot 98: 572-592.

Accepted Article

517

21 This article is protected by copyright. All rights reserved.

Caron, D.A., Countway, P.D., Savai, P., et al. (2009) Defining DNA-based operational

544

taxonomic units for microbial-eukaryote ecology. Appl Environ Microbiol 75: 5797-5808.

545

Chan, A.T. (1978) Comparative physiological study of marine diatoms and dinoflagellates in

546

relation to irradiance and cell size. I. Growth under continuous light. J Phycol 14: 396-402.

547

Charvet, S., Vincent, W., Comeau, A. and Lovejoy, C. (2012) Pyrosequencing analysis of the

548

protist communities in a High Arctic meromictic lake: DNA preservation and change. Front

549

Microbiol 3: 422.

550

Cloern, J.E., Cole, B.E. and Oremland, R.S. (1983) Seasonal changes in the chemistry and

551

biology of a meromictic lake (Big Soda Lake, Nevada, U.S.A.). Hydrobiologia 105: 195-206.

552

Csárdi, G. and Nepusz, T. (2006) The Igraph Software Package for Complex Network

553

research. Interjournal Complex Syst Manuscript number: 1695.

554

Dunthorn, M., Klier, J., Bunge, J. and Stoeck, T. (2012) Comparing the hyper-variable V4

555

and V9 regions of the Small Subunit rDNA for assessment of ciliate environmental diversity.

556

J Eukaryot Microbiol 59: 185-187.

557

Dunthorn, M., Stoeck, T., Clamp, J., Warren, A. and Mahé, F. (2014) Ciliates and the Rare

558

Biosphere: A Review. J Eukaryot Microbiol 61: 404-409.

559

Edgar, R.C. (2010) Search and clustering orders of magnitude faster than BLAST.

560

Bioinformatics 26: 2460-2461

561

Edgar, R.C., Haas, B.J., Clemente, J.C., Quince, C. and Knight, R. (2011) UCHIME

562

improves sensitivity and speed of chimera detection. Bioinformatics 27: 2194-2200.

563

Ehrenberg, C.C. (1838) Die Infusionstheirchen als Vollkommene Organismen, Leopold

564

Voss, Leipzig, Germany.

565

Esteban, G.F., Finlay, B.J. and Clarke, K.J. (2012) Priest Pot in the English Lake District: a

566

showcase of microbial diversity. Freshw Biol 57: 321-330.

567

Fenchel, T. and Finlay, B.J. (1995) Ecology and Evolution in Anoxic Worlds.

568

New York, NY, USA: Oxford University Press

Accepted Article

543

22 This article is protected by copyright. All rights reserved.

Fenchel, T. and Finlay, B. (2008) Oxygen and the spatial structure of microbial communities.

570

Biol Rev 83: 553-569.

571

Fenchel, T., Kristensen, L.D. and Rasmussen, L. (1990) Water column anoxia: vertical

572

zonation of planktonic protozoa. Mar Ecol Prog Ser 62: 1-10.

573

Finlay, B.J. and Esteban, G.F. (1998) Freshwater protozoa: biodiversity and ecological

574

function. Biodiversity Conserv 7: 1163-1186.

575

Finlay, B.J., Maberly, S.C. and Esteban, G.F. (1996) Spectacular abundance of ciliates in

576

anoxic pond water: contribution of symbiont photosynthesis to host respiratory oxygen

577

requirements. FEMS Microbiol Ecol 20: 229-235.

578

Flaim, G., Obertegger, U. and Guella, G. (2012) Changes in galactolipid composition of the

579

cold freshwater dinoflagellate Borghiella dodgei in response to temperature. Hydrobiologia

580

698: 285-293.

581

Foissner, W., Chao, A. and Katz, L.A. (2008) Diversity and geographic distribution of

582

ciliates (Protista: Ciliophora). Biodiversity Conserv 17: 345-363.

583

Forster, D., Behnke, A. and Stoeck, T. (2012) Meta-analyses of environmental sequence data

584

identify anoxia and salinity as parameters shaping ciliate communities. System Biodivers 10:

585

277-288.

586

Fritz, G., Pfannkuchen, M., Struck, U., Hengherr, S., Strohmeier, S. and Brümmer, F. (2012)

587

Characterizing an Anoxic Habitat: Sulfur Bacteria in a Meromictic Alpine Lake. In Anoxia.

588

Altenbach, A.V., Bernhard, J.M. and Seckbach, J. (eds). Netherlands: Springer, pp. 449-461.

589

Fröbisch, G., Mangelsdorf, J., Schauer, T., Streil, J. and Wachter, H. (1977)

590

Gewässerkundliche Studie über sechs Seen bei Füssen im Allgäu. Informations des Bayer

591

Landesamtes für Wasserwirtschaft, München 3/77 Munich.

592

Galtier, N., Gouy, M. and Gautier, C. (1996) SEAVIEW and PHYLO_WIN: two graphic

593

tools for sequence alignment and molecular phylogeny. Comput Appl Biosci 12: 543-548.

Accepted Article

569

23 This article is protected by copyright. All rights reserved.

Gasol, J.M., García-Cantizano, J., Massana, R., Guerrero, R. and Pedrós-Alió, C. (1993)

595

Physiological ecology of a metalimnetic Cryptomonas population: relationships to light,

596

sulfide and nutrients. J Plankton Res 15: 255-275.

597

Gervais, F. (1998) Ecology of cryptophytes coexisting near a freshwater chemocline. Freshw

598

Biol 39: 61-78.

599

Glücksman, E., Bell, T., Griffiths, R.I. and Bass, D. (2010) Closely related protist strains

600

have different grazing impacts on natural bacterial communities. Environ Microbiol 12: 3105-

601

3113.

602

Gobet, A., Quince, C. and Ramette, A. (2010) Multivariate Cutoff Level Analysis

603

(MultiCoLA) of large community data sets. Nucleic Acids Res 38: e155.

604

Guhl, B.E., Finlay, B.J. and Schink, B. (1996) Comparison of ciliate communities in the

605

anoxic hypolimnia of three lakes: general features and the influence of lake characteristics. J

606

Plankton Res 18: 335-353.

607

Haegeman, B., Hamelin, J., Moriarty, J., Neal, P., Dushoff, J. and Weitz, J. (2013) Robust

608

estimation of microbial diversity in theory and in practice. ISME J 7: 1092-1101.

609

Jakobsen, H., Hansen, P.J. and Larsen, J. (2000) Growth and grazing responses of two

610

chloroplast-retaining dinoflagellates : effect of irradiance and prey species. Mar Ecol Prog

611

Ser 201: 121-128.

612

James, M.R., Burns, C.W. and Forsyth, D.J. (1995) Pelagic ciliated protozoa in two

613

monomictic, southern temperate lakes of contrasting trophic state: seasonal distribution and

614

abundance. J Plankton Res 17: 1479-1500.

615

Jürgens, K. and Massana, R. (2008) Protistan grazing on marine bacterioplankton. In

616

Microbial Ecology of the Oceans. Kirchman, L. (ed). New York: Wiley-Liss, pp. 383-441

617

Klaveness, D. and Løvhøiden, F. (2007) Meromictic lakes as habitats for protists. In Algae

618

and Cyanobacteria in Extreme Environments. Seckbach, J. (ed). Netherlands: Springer, pp.

619

59-78.

Accepted Article

594

24 This article is protected by copyright. All rights reserved.

Kolodziej, K. and Stoeck, T. (2007) Cellular identification of a novel uncultured Marine

621

Stramenopile (MAST-12 Clade) Small-Subunit rRNA gene sequence from a Norwegian

622

estuary by use of Fluorescence In Situ Hybridization-Scanning Electron Microscopy. Appl

623

Environ Microbiol 73: 2718-2726.

624

Lane, D.J. (1991) 16S / 23S sequencing. In Nucleic Acid Technologies in Bacterial

625

Systematic. Stackebrandt, E. and Goodfellow, M. (eds). New York: Wiley, pp. 115-175.

626

Lefèvre, E., Bardot, C., Noël, C., Carrias, J.F., Viscogliosi, E., Amblard, C. and Sime-

627

Ngando, T. (2007) Unveiling fungal zooflagellates as members of freshwater picoeukaryotes:

628

evidence from a molecular diversity study in a deep meromictic lake. Environ Microbiol 9:

629

61-71.

630

Lepère, C., Boucher, D., Jardillier, L., Domaizon, I. and Debroas, D. (2006) Succession and

631

regulation factors of small eukaryote community composition in a lacustrine ecosystem (Lake

632

Pavin). Appl Environ Microbiol 72: 2971-2981.

633

Lindström, K. (1984) Effect of temperature, ligh and pH on growth, photosynthesis and

634

respiration of the dinoflagellate Peridinium cinctum FA. WestII in laboratory cultures. J

635

Phycol 20: 212-220.

636

Logares, R., Bråte, J., Bertilsson, S., Clasen, J.L., Shalchian-Tabrizi, K. and Rengefors, K.

637

(2009) Infrequent marine-freshwater transitions in the microbial world. Trends Microbiol 17:

638

414-422.

639

Luo, Q., Krumholz, L.R., Najar, F.Z., Peacock, A.D., Roe, B.A., White, D.C. and Elshahed

640

MS (2005) Diversity of the microeukaryotic community in sulfide-rich Zodletone spring

641

(Oklahoma). Appl Environ Microbiol 71: 6175-6184.

642

Luo, W., Bock, C., Li, H.R., Padisák, J. and Krienitz, L. (2011) Molecular and microscopic

643

diversity of planktonic eukaryotes in the oligotrophic Lake Stechlin (Germany).

644

Hydrobiologia 661: 133-143.

Accepted Article

620

25 This article is protected by copyright. All rights reserved.

Lüthy, L., Fritz, M. and Bachofen, R. (2000) In situ determination of sulfide turnover rates in

646

a meromictic alpine lake. Appl Environ Microbiol 66: 712-717.

647

Lynch, M.D.J., Bartram, A.K. and Neufeld, J.D. (2012) Targeted recovery of novel

648

phylogenetic diversity from next-generation sequence data. ISME J 6: 2067-2077.

649

Lynn, D.H. and Small, E.B. (2002) Phylum Ciliophora. In An illustrated guide to the

650

Protozoa. Lee, J.J., Bradbury, P.C. and Leedale, G.F. (eds). KS: Society of Protozoologists,

651

Lawrence, pp. 371-656.

652

Magoč, T. and Salzberg, S.L. (2011) FLASH: Fast Length Adjustment of Short Reads to

653

improve genome assemblies. Bioinformatics 27: 2957-2963.

654

Mahé, F., Rognes, T., Quince, C., de Vargas, C. and Dunthorn, M. (2014) Swarm: robust and

655

fast clustering method for amplicon-based studies. PeerJ 2:e593; DOI 10.7717/peerj.593

656

Maruyama, S. and Kim, E. (2013) A modern descendant of early green algal phagotrophs.

657

Curr Biol 23: 1081-1084.

658

Massana, R. and Pedrós-Alió, C. (2008) Unveiling new microbial eukaryotes in the surface

659

ocean. Curr Opin Microbiol 11: 213-218.

660

Massana, R., del Campo, J., Sieracki, M.E., Audic, S. and Logares, R. (2013) Exploring the

661

uncultured microeukaryote majority in the oceans: reevaluation of ribogroups within

662

stramenopiles. ISME J 8: 854-866.

663

Medlin, L., Elwood, H., Stickel, S. and Sogin, M. (1988) The characterization of

664

enzymatically amplified eukaryotic 16S-like rRNA-coding regions. Gene 71: 491-499.

665

Meuser, J.E., Ananyev, G., Wittig, L.E., et al. (2009) Phenotypic diversity of hydrogen

666

production in chlorophycean algae reflects distinct anaerobic metabolisms. J Biotechnol 142:

667

21-30.

668

Miracle, M.R., Vicente, E. and Pedrós-Alió, C. (1992) Biological studies of spanish

669

meromictic and stratified karstic lakes. Limnetica 8: 59-77.

Accepted Article

645

26 This article is protected by copyright. All rights reserved.

Nebel, M.E., Wild, S., Holzhauser, M., Hüttenberger, L., Reitzig, R., Sperber, M. and Stoeck,

671

T. (2011) JAGUC - A software package for environmental diversity analyses. J Bioinform

672

Comput Biol 09: 749-773.

673

Oikonomou, A., Pachiadaki, M. and Stoeck, T. (2014) Protistan grazing in a meromictic

674

freshwater lake with anoxic bottom water. FEMS Microbiol Ecol 87: 691-703.

675

Orsi, W., Edgcomb, V., Jeon., S., et al. (2011) Protistan microbial observatory in the Cariaco

676

Basin, Caribbean. II. Habitat specialization. ISME J 5: 1357-1373.

677

Park, J.S. and Simpson, A.G.B. (2010) Characterization of halotolerant Bicosoecida and

678

Placididea (Stramenopila) that are distinct from marine forms, and the phylogenetic pattern of

679

salinity preference in heterotrophic stramenopiles. Environ Microbiol 12: 1173-1184.

680

Pawlowski, J. and Holzmann, M. (2014) A plea for DNA barcoding of Foraminifera. The J

681

Foraminiferal Res 44: 62-67.

682

Pedrós-Alió, C., Massana, R., Latasa, M., Garcí-Cantizano, J. and Gasol, J.M. (1995)

683

Predation by ciliates on a metalimnetic Cryptomonas population: feeding rates, impact and

684

effects of vertical migration. J Plankton Res 17: 2131-2154.

685

Pfister, G., Auer, B. and Arndt, H. (2002) Pelagic ciliates (Protozoa, Ciliophora) of different

686

brackish and freshwater lakes - a community analysis at the species level. Limnologica 32:

687

147-168.

688

Pimenov, N.V., Rusanov, I.I., Karnachuk, O.V., et al. (2003) Microbial processes of the

689

carbon and sulfur cycles in Lake Shira (Khakasia). Microbiology 72: 221-229.

690

Pommier, T., Neal, P.R., Gasol, J.M., Coll, M., Acinas, S.G. and Pedrós-Alió, C. (2010)

691

Spatial patterns of bacterial richness and evenness in the NW Mediterranean Sea explored by

692

pyrosequencing of the 16S rRNA. Aquat Microb Ecol 61: 221-233.

693

Prézelin, B.B. and Matlick, H.A. (1983) Nutrient-dependent low-light adaptation in the

694

dinoflagellate Gonyaulax polyedra. Mar Biol 74: 141-150.

Accepted Article

670

27 This article is protected by copyright. All rights reserved.

Richards, T.A., Vepritskiy, A.A., Gouliamova, D.E. and Nierzwicki-Bauer, S.A. (2005) The

696

molecular diversity of freshwater picoeukaryotes from an oligotrophic lake reveals diverse,

697

distinctive and globally dispersed lineages. Environ Microbiol 7: 1413-1425.

698

Roberts, E.C., Legrand, C., Steinke, M. and Wootton, E.C. (2011) Mechanisms underlying

699

chemical interactions between predatory planktonic protists and their prey. J Plankton Res 33:

700

833-841.

701

Saccà, A., Guglielmo, L. and Bruni, V. (2008) Vertical and temporal microbial community

702

patterns in a meromictic coastal lake influenced by the Straits of Messina upwelling system.

703

Hydrobiologia 600: 89-104.

704

Sanders, R.W. (2011) Alternative nutritional strategies in protists: Symposium introduction

705

and a review of freshwater protists that combine photosynthesis and heterotrophy. J Eukaryot

706

Microbiol 58: 181-184.

707

Shendure, J. and Ji, H. (2008) Next-generation DNA sequencing. Nat Biotech 26: 1135-1145.

708

Sherr, E.B. and Sherr, B.F. (2002) Significance of predation by protists in aquatic microbial

709

food webs. Antonie van Leeuwenhoek 81: 293-308.

710

Šimek, K., Kasalicky, V., Jezbera, J., et al. (2013) Differential freshwater flagellate

711

community response to bacterial food quality with a focus on Limnohabitans bacteria. ISME J

712

7: 1519-1530.

713

Šlapeta, J., Moreira, D. and López-García, P. (2005) The extent of protist diversity: insights

714

from molecular ecology of freshwater eukaryotes. Proc Biol Sci 272: 2073-2081.

715

Smayda, T.J. (2002) Adaptive ecology, growth strategies and the global bloom expansion of

716

dinoflagellates. J Oceanogr 58: 281-294.

717

Stewart, K., Walker, K.F. and Likens, G.E. (2009) Meromictic lakes. In Encyclopedia of

718

Inland Waters. Likens, G.E. (ed) Oxford: Elsevier, pp.589-602.

Accepted Article

695

28 This article is protected by copyright. All rights reserved.

Stief, P., Fuchs-Ocklenburg, S., Kamp, A., et al. (2014) Dissimilatory nitrate reduction by

720

Aspergillus terreus isolated from the seasonal oxygen minimum zone in the Arabian Sea.

721

BMC Microbiol 14: 35.

722

Stock, A., Jürgens, K., Bunge, J. and Stoeck, T. (2009) Protistan diversity in suboxic and

723

anoxic waters of the Gotland Deep (Baltic Sea) as revealed by 18S rRNA clone libraries.

724

Aquat Microb Ecol 55: 267-284.

725

Stoeck, T., Taylor, G.T. and Epstein, S.S. (2003) Novel eukaryotes from the permanently

726

anoxic Cariaco Basin (Caribbean Sea). Appl Environ Microbiol 69: 5656-5663.

727

Stoeck, T., Zuendorf, A., Breiner, H.W. and Behnke, A. (2007) A molecular approach to

728

identify active microbes in environmental eukaryote clone libraries. Microb Ecol 53: 328-339.

729

Stoeck, T., Hayward, B., Taylor, G.T., Varela, R. and Epstein, S.S. (2006) A multiple PCR-

730

primer approach to access the microeukaryotic diversity in environmental samples. Protist

731

157: 31-43.

732

Stoeck, T., Breiner, H.W., Filker, S., Ostermaier, V., Kammerlander, B. and Sonntag, B.

733

(2014) A morphogenetic survey on ciliate plankton from a mountain lake pinpoints the

734

necessity of lineage-specific barcode markers in microbial ecology. Environ Microbiol 16:

735

430-444.

736

Stoeck, T., Bass, D., Nebel, M., Christen, R., Jones, M.D.M., Breiner, H.W. and Richards,

737

T.A. (2010) Multiple marker parallel tag environmental DNA sequencing reveals a highly

738

complex eukaryotic community in marine anoxic water. Mol Ecol 19: 21-31.

739

Storelli, N., Peduzzi, S., Saad, M.M., Frigaard, N-U., Perret, X. and Tonolla, M. (2013) CO2

740

assimilation in the chemocline of Lake Cadagno is dominated by a few types of phototrophic

741

purple sulfur bacteria. FEMS Microbiol Ecol 84: 421-432.

742

Tarbe, A.L., Stenuite, S., Balagué, V., Sinyinza, D., Descy, J.P. and Massana, R. (2011)

743

Molecular characterisation of the small-eukaryote community in a tropical Great Lake (Lake

744

Tanganyika, East Africa). Aquat Microb Ecol 62: 177-190.

Accepted Article

719

29 This article is protected by copyright. All rights reserved.

Taylor, F.J.R., Hoppenrath, M. and Saldarriaga, J. (2008) Dinoflagellate diversity and

746

distribution. Biodiversity Conserv 17: 407-418.

747

Triadó-Margarit, X. and Casamayor, E.O. (2012) Genetic diversity of planktonic eukaryotes

748

in high mountain lakes (Central Pyrenees, Spain). Environ Microbiol 14: 2445-2456.

749

Tuomi, P., Torsvik, T., Heldal, M. and Bratbak, G. (1997) Bacterial population dynamics in a

750

meromictic lake. Appl Environ Microbiol 63: 2181-2188.

751

Weis, W. (1983) Die postglaziale Ablagerungsgeschichte des Alpsees und Alatsee,

752

Südbayern. Thesis (Diploma), Ludwig Maximilians University, Institute of Geography,

753

Munich.

754

Wylezich, C. and Jürgens, K. (2011) Protist diversity in suboxic and sulfidic waters of the

755

Black Sea. Environ Microbiol 13: 2939-2956.

756

Zingel, P. (2005) Vertical and seasonal dynamics of planktonic ciliates in a strongly stratified

757

hypertrophic lake. Hydrobiologia 547: 163-174.

758

Zingel, P. and Ott, I. (2000) Vertical distribution of planktonic ciliates in strongly stratified

759

temperate lakes. Hydrobiologia 435: 19-26.

760

Zinger, L., Gobet, A. and Pommier, T. (2012) Two decades of describing the unseen majority

761

of aquatic microbial diversity. Mol Ecol 21: 1878-1896.

Accepted Article

745

762 763

Table and Figure legends

764 765

Fig. 1. Depth profiles of oxygen, temperature, and sulfide in (A) autumn 2011, (B) spring

766

2012 and (C) summer 2012 in meromictic lake Alatsee. The grey area indicates the transition

767

zone.

768 769

Fig. 2. Number of OTUs as a function of clustering threshold.

770 30 This article is protected by copyright. All rights reserved.

Fig. 3. Alpha diversity within each particular sample defined by the Simpson index (1-D).

772

Samples correspond to MIXO: mixolimnion (oxic), IF: interface, MONO: monimolimnion

773

(anoxic) and Au: autumn, Sp: spring, Su: summer.

Accepted Article

771

774 775

Fig. 4. Hierarchical clustering (Jabundance) of samples based on protistan and fungal OTUs95%,

776

after excluding OTUs that include less than three sequences. Bootstrap values indicate the

777

strength of the relationships in the UPGMA dendrogramm. Samples correspond to MIXO:

778

mixolimnion (oxic), IF: interface, MONO: monimolimnion (anoxic) and Au: autumn, Sp:

779

spring, Su: summer. The length of the reference bar represents a linkage distance of 0.02.

780 781

Fig. 5. Relative taxonomic distribution of classified protistan and fungal OTUs95%.

782

Taxonomic groups that were represented by a proportion of ≤ 1% of total number of OTUs in

783

at least one of the samples are grouped in the category Others. Samples correspond to MIXO:

784

mixolimnion (oxic), IF: interface, MONO: monimolimnion (anoxic) and Au: autumn, Sp:

785

spring, Su: summer.

786 787

Fig. 6. Taxonomic distribution of OTUs found exclusively in (A) oxic mixolimnion, (B)

788

interface and (C) anoxic monimolimnion in all examined seasons. Underlined color code

789

corresponds to inner circles. Outer rings represent the subcategories of the inner circles with

790

> 5% of total number of OTUs and correspond to the numbered colors. Number (#) of OTUs

791

denotes the total number of OTUs found exclusively in all seasons in each water mass. The

792

category “Others” denotes groups that were represented by a proportion of ≤ 1% of total

793

number of OTUs in each water mass.

794 795

Fig. 7. Divergent genetic diversity network of 200 most abundant swarms (OTUs95%) in lake

796

Alatsee, pooled over depth and seasons. The node size is indicative of the number of 31 This article is protected by copyright. All rights reserved.

sequences in each swarm. Two nodes are connected if they share a sequence similarity of at

798

least 90%. The novelty level is denoted by BLAST identity against Genbank sequences, red-

799

orange: highly divergent swarms.

Accepted Article

797

800 801

Fig. S1. Hierarchical clustering (Jbinary) of samples based on protistan and fungal OTUs95%,

802

after excluding OTUs that include less than three sequences. Bootstrap values indicate the

803

strength of the relationships in the UPGMA dendrogramm. Samples correspond to MIXO:

804

mixolimnion (oxic), IF: interface, MONO: monimolimnion (anoxic) and Au: autumn, Sp:

805

spring, Su: summer. Protistan and fungal OTUs were clustered at a 95% similarity level. Bar

806

scales represent the proportion of linkage distance.

807 808

Table S1. Hexamer identifiers (in bold) used for distinguishing between the samples. Pr:

809

denotes the position of either the forward 1391F or the reverse EukB primer. We considered

810

carefully the combination of a highly conserved forward primer and a eukaryotic-specific

811

reverse primer to avoid biases towards few eukaryotic taxa. The initial bases (underlined)

812

were added before the hexamers to avoid a too strong light signal generated by the first bases

813

at the beginning of the sequencing procedure. Samples 5´-Initial bases-Hexamer-Pr-3´ MIXO-Au TGC-AGCGTC-Pr IF-Au TAGT-ATACTC-Pr MONO-Au C-ATCGAC-Pr MIXO-Sp T-ACACAC-Pr IF-Sp AT-ACGAGC-Pr MONO-Sp ATC-AGACGC-Pr MIXO-Su ACAG-ACTATC-Pr IF-Su T-AGCTAC-Pr MONO-Su TC-ATACTC-Pr

814

32 This article is protected by copyright. All rights reserved.

Oxygen concentration [mg l-1] & Temperature [°C]

A

0

5

10

15

20

25

80

100

Oxygen 5

Temperature Sulfide

10

Depth

Accepted Article

0

15 20 25 30 0

20

40

60

Sulfide concentration [mg l-1]

B

Oxygen concentration [mg l-1] & Temperature [°C] 0

5

10

15

20

25

0

20

40

60

80

100

0

5

Depth

10

15

20

25

30

Sulfide concentration [mg

l-1]

Oxygen concentration [mg l-1] & Temperature [°C]

C

0

5

0

20

10

15

20

25

40

60

80

100

0

5

Depth

10

15

20

25

30

EMI_12666_F1.eps

Sulfide concentration [mg l-1]

Accepted Article 600000

Number of OTUs

500000

400000

300000

200000

100000

0

100

EMI_12666_F2.eps

99

98

97

96

95

94

Clustering threshold (%)

93

92

91

90

89

Accepted Article 0.99 0.97

Simpson index

0.95 0.93 0.91 0.89 0.87 0.85

EMI_12666_F3.eps

Samples

Accepted Article

EMI_12666_F4.eps

MIXO-Au

1

MIXO-Sp

0.88

MIXO-Su IF-Au

0.98

IF-Su 1

1 0.98 0.98

IF-Sp MONO-Sp MONO-Au MONO-Su

Accepted Article 100% 90%

Percentage of classified OTUs

80%

Others

70%

Stramenopiles Metamonada

60%

Amoebozoa Euglenozoa

50%

Fungi

40%

Chlorophyta Cryptophyta

30%

Dinoflagellata

Ciliophora

20% 10%

0%

EMI_12666_F5.eps

Samples

Accepted Article A)

B)

MIXOLIMNION # of OTUs: 412

1

3

C)

INTERFACE

# of OTUs: 23

1

1

2 4

7 8 3 6

2

5 2

3

1

6 3

3

3

2

1

1

2

1

1

# of OTUs: 111

1

2

2

MONIMOLIMNION

1

2 3 1

2 9 6 43

5

1 6

4

2 5

1

3

Ciliophora

Stramenopiles

Euglenozoa

Metamonada

Centrohelida

1 2 3 4 5 6 7 8 9

1 2 3 4 5 6

1 Kinetoplastea 2 Euglinida

1 Diplomonadida 2 Preaxostyla 3 Parabasalia Chlorophyta

Protalveolata

1 Ulvophyceae 2 Chlorophyceae 3 Mamiellophyceae Cercozoa

Others

Spirotrichea Colpodea Protostomatea Nassophorea Heterotricheaa Oligohymenophorea Armophorea Karyorelictea Other Ciliophora

EMI_12666_F6.eps

Chrysophyceae Synurales Raphidiophyceae Bolidomonas Bicosoecida Other Stramenopiles

Cryptophyta 1 Cryptomonadales 2 Pyrenomonadales

Dinoflagellata

Amoebozoa

1 Peridiniphycidae 2 Gymnodiniphycidae 3 Other Dinoflagellata

1 Archamoebae 2 Discocea 3 Other Amoebozoa

1 Cercomonadidae 2 Other Cercozoa

Fungi Apusomonadida

Accepted Article

EMI_12666_F7.eps

Protistan diversity in a permanently stratified meromictic lake (Lake Alatsee, SW Germany).

Protists play a crucial role for ecosystem function(ing) and oxygen is one of the strongest barriers against their local dispersal. However, protistan...
804KB Sizes 1 Downloads 18 Views