Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Surprising Prokaryotic and Eukaryotic Diversity, Community Structure and Biogeography of Ethiopian Soda Lakes

  • Anders Lanzén ,

    alanzen@neiker.net

    Current address: Department of Ecology and Natural Resources, NEIKER-Tecnalia, Derio, Spain

    Affiliations Department of Biology and Centre for Geobiology, University of Bergen, Bergen, Norway, Computational Biology Unit, Uni Computing, Uni Research AS, Bergen, Norway

  • Addis Simachew,

    Affiliation Institute of Biotechnology, Addis Ababa University, Addis Ababa, Ethiopia

  • Amare Gessesse,

    Affiliation Institute of Biotechnology, Addis Ababa University, Addis Ababa, Ethiopia

  • Dominika Chmolowska,

    Affiliation Institute of Environmental Sciences, Jagiellonian University, Krakow, Poland

  • Inge Jonassen,

    Affiliations Computational Biology Unit, Uni Computing, Uni Research AS, Bergen, Norway, Department of Informatics, University of Bergen, Bergen, Norway

  • Lise Øvreås

    Affiliation Department of Biology and Centre for Geobiology, University of Bergen, Bergen, Norway

Abstract

Soda lakes are intriguing ecosystems harboring extremely productive microbial communities in spite of their extreme environmental conditions. This makes them valuable model systems for studying the connection between community structure and abiotic parameters such as pH and salinity. For the first time, we apply high-throughput sequencing to accurately estimate phylogenetic richness and composition in five soda lakes, located in the Ethiopian Rift Valley. The lakes were selected for their contrasting pH, salinities and stratification and several depths or spatial positions were covered in each lake. DNA was extracted and analyzed from all lakes at various depths and RNA extracted from two of the lakes, analyzed using both amplicon- and shotgun sequencing. We reveal a surprisingly high biodiversity in all of the studied lakes, similar to that of freshwater lakes. Interestingly, diversity appeared uncorrelated or positively correlated to pH and salinity, with the most “extreme” lakes showing the highest richness. Together, pH, dissolved oxygen, sodium- and potassium concentration explained approximately 30% of the compositional variation between samples. A diversity of prokaryotic and eukaryotic taxa could be identified, including several putatively involved in carbon-, sulfur- or nitrogen cycling. Key processes like methane oxidation, ammonia oxidation and ‘nitrifier denitrification’ were also confirmed by mRNA transcript analyses.

Introduction

Soda lakes are strongly alkaline lakes, typically with a pH between 9 to 11, high concentrations of carbonate ions and with salinities ranging from brackish to hypersaline [1]. Although relatively rare, these lakes constitute a large part of inland water by volume in certain regions, particularly arid or semi-arid areas connected to tectonic rifts such as the East African Rift Valley. In spite of their basicity, many soda lakes show unusually high primary productivity, including the highest photosynthesis rates measured in any aquatic habitat (above 7 g C m−2 day−1) [2]. Thus, they rank not only as the most productive but also among the most extreme aquatic ecosystems. Not all soda lakes are highly productive, however, and the mechanisms controlling primary production may involve many factors such as nutrient limitations, toxicity, or trophic interactions [3]. Salinity, however, may be the strongest stress factor limiting microbial diversity [4], [5]. In spite of this, high morphological diversity comparable to neutral freshwater systems has been observed, even in hypersaline soda lakes [3].

In this study, we address the diversity of these fascinating ecosystems, challenging the notion that extreme habitats generally harbor lower biodiversity. Further, we investigate whether a relationship between salinity and taxonomic richness exists in the investigated soda lakes. The underlying question we attempt to answer is to what extent salinity, pH and other parameters influence the microbial community structure. We also address whether lakes located closer together generally harbored more similar communities.

A number of alkaliphilic microorganisms with key metabolic roles in soda lakes have previously been studied using both cultivation dependent and independent techniques. Such studies have uncovered a diversity of various functional and phylogenetic groups including cyanobacteria (e.g. [6]), anoxygenic phototrophs (e.g. [7], [8], [9], [10]), aerobic heterotrophs (e.g. [10]), sulfur reducers and other anaerobic organotrophs (e.g. [3]), sulfur oxidizers (reviewed in [11]), acetogens (e.g. [5]), methanogens (e.g. [12]), methylotrophs (e.g. [13], [14], [15]) or eukaryotic microorganisms (e.g. [16]). Others addressed planktonic community composition and diversity of individual lakes [17], [18] or across several lakes, including soda lakes [19], [20], by using molecular fingerprinting or SSU rRNA clone libraries. However, such techniques are limited in terms of their resolution or ability to determine the phylogenetic diversity and community structure at high resolution [21].

This study is the first to use high throughput sequencing to analyze the structure of soda lake microbial plankton communities. Using a combination of traditional marker gene profiling and PCR-independent shotgun sequencing of reverse transcribed rRNA, we target both the active (RNA) and present (DNA) diversity and composition in these intriguing ecosystems. This approach has been shown to provide a more holistic view [22], [23], enabling quantitative analysis of taxonomic groups from all domains of life simultaneously as well as a snapshot of abundant mRNA transcripts.

Results

Characteristics of Investigated Lakes

Five soda lakes were selected based on contrasting characteristics such as salinity, size and location. Lakes Abijata, Chitu and Shalla are located more centrally within the East African Rift and share higher salinities compared to Beseka and Arenguadi, located in the Upper Awash Basin and Central Ethiopian Highlands, respectively. While Arenguadi and Chitu are relatively small crater lakes (the former a maar), Shalla is the largest studied and represent the only deep lake, with maximum and average depths of 266 and 87 m, respectively [24]. While Abijata is retracting due to water diversion and soda ash extraction [25], Beseka is expanding [26], resulting in changing salinities and phytoplankton communities [27]. Tables 1 and S1 list the sampling sites, collected samples and physicochemical parameters measured.

thumbnail
Table 1. Overview of the soda lakes, samples and sequence datasets studied.

https://doi.org/10.1371/journal.pone.0072577.t001

Significant stratification was only encountered in Lake Arenguadi, saturated in oxygen until a depth of 3 m, followed by a narrow oxycline and then anoxia below 4 m. Subtle changes in salinity and pH were also encountered between limnia (Table S1). The holomictic lake Chitu appeared recently mixed during sampling. Only low levels of oxygen were measured at water surface until about 10 cm depth with no discernible salinity or pH gradients.

Composition and Diversity of the Microbial Communities

In total 458,813 sequence reads representing SSU rRNA were obtained from DNA amplicon- (n = 22), cDNA amplicon- (n = 8) and shotgun sequence (n = 2) datasets, in addition to 6,745 putative mRNA reads (Tables 1 and S2). The “prefilter” samples from lakes Beseka and Chitu, yielded disproportionately large and small numbers of sequence reads, respectively.

Total OTU richness amounted to 2,704 (3% distance), excluding 1,286 singletons. OTUs per dataset varied between 169 and 1,519 (Table S2). As indicated by rarefaction analysis (Fig. S1), sequencing depth was far from being exhaustive even in the largest dataset. Estimated Shannon diversity (H′) varied between 2.3 and 4.7, showing no correlation to the number of reads, as opposed to OTU richness (Table S2). However, the substantial variance of H′ between spatial replicates inside of Abijata and Chitu was similar to variance between lakes, indicating that differences in H’ between lakes were not significant, at least lacking better replication. Instead, Bayesian parametric estimation of total richness [28] was used to compare diversity between datasets in a more accurate manner. The Sichel distribution fit best to the observed prokaryotic OTU-abundance distributions in most datasets (28 of 30) and was used to calculate confidence intervals of total sample richness, illustrated in Figure 1. Medians of estimated richness ranges generally followed the same trend as rarefied OTU richness, but the later varied more across spatial replicates, consistent with the variance of H′.

thumbnail
Figure 1. Parametric richness estimates.

Box-plots cover 95% Bayesian confidence intervals of total OTU richness for each sample. Grey boxes indicate DNA amplicon datasets, white boxes cDNA amplicons and black boxes DNA amplicon datasets derived from prefilters. Solid lines below the box plots indicate rarified OTU richness. Arithmetic means of medians for DNA amplicon datasets (excluding prefilter-derived) are shown below lake names.

https://doi.org/10.1371/journal.pone.0072577.g001

The highest median richness was estimated in the surface of Beseka and the lowest at 30 m depth in Shalla. Significantly higher richness (α = 0.05) was estimated from several datasets compared to the later. From means of estimated richness in DNA datasets (excluding prefilters; Fig. 1), Chitu appears to be the most diverse lake, closely followed by Abijata. The RNA-derived datasets showed similar richness estimates as their corresponding DNA datasets and followed the same trend, except in two cases (Fig. 1). Firstly, estimated RNA richness was significantly lower in the Beseka surface sample. Secondly, in Arenguadi at 2 m, significantly higher richness was predicted in RNA. Predicted richness in the stratified Lake Arenguadi followed a trend remarkably similar to that of cell density, as estimated using DAPI staining (Fig. S2).

Comparisons of Community Structure and Influence of Physicochemical Parameters

Figure 2 shows the distribution of OTUs across lakes (excluding prefilter- and cDNA-derived plus adjusted for contrasting sequencing depths). Abijata and Shalla showed a relatively larger overlap than other lakes, while Beseka harbored most OTUs unique to one lake. Larger proportions of OTUs were shared between different depths in the same lake, compared to those shared between lakes, particularly for RNA-derived datasets (Fig. S3).

thumbnail
Figure 2. Venn diagram showing the distribution of shared OTUs across lakes.

White numbers indicate the number of OTUs in each possible subset, adjusted for differences in sequencing depth.

https://doi.org/10.1371/journal.pone.0072577.g002

Hierarchical clustering (Fig. S4) and Non-metric multidimensional scaling (NMDS; Fig. 3) based on OTU composition (Bray-Curtis dissimilarities) showed that all datasets formed lake-specific clusters, except for Arenguadi where the anoxic hypolimnion (10 and 30 m, “Arenguadi deep”) formed a separate cluster from the oxic epilimnion (0–3 m, “Arenguadi shallow”). The former appeared more similar to Chitu, representing the other anoxic environment; and the later to Beseka, representing the other low-salinity lake. The same clustering pattern was obtained using taxonomical distributions rather than OTUs, with the two shotgun-sequenced datasets clustering with their respective lakes (Fig. S5). Based on the observed clustering pattern, six “habitat” clusters were defined.

thumbnail
Figure 3. Non-metric multidimensional scaling (NMDS) based on Bray-Curtis dissimilarities between OTU compositions of individual datasets.

Sequence datasets OTUs and fitted physicochemical parameters are plotted on the first two NMDS axes. The colors and shapes of individual OTUs and sequence datasets represent their taxonomical classification or dataset type, according to the legends.

https://doi.org/10.1371/journal.pone.0072577.g003

As indicated by NMDS, community samples within lakes were more similar than those from different lakes, with the exception of the two layers found in Arenguadi. This pattern was confirmed comparing dissimilarities (Bray-Curtis) between shallow samples inside the same lakes (for Arenguadi and Chitu), to those between different lakes (using average compositions for replicate samples). According to a Welch t-test the difference in similarity was significant (p<10−15).

In order to evaluate the influence of lake water composition and other measured physicochemical parameters, a separate NMDS was constructed from pooled habitat datasets. Out of the parameters, four showed significant correlation to this NMDS: dissolved oxygen (presence or absence), pH, sodium- (Na+) and potassium (K+) concentrations. These parameters also correlated significantly to the NMDS made from un-merged datasets and their fitted vectors have been added in Fig. 3. Variation partioning analysis suggested that taken together, these parameters explained 29% of the variation in community composition between habitats and 31% between individual datasets (Fig. S6). The influence of distance between lakes on community dissimilarity was also investigated using linear regression (Fig. S7). A weak correlation may exist, but was not significant among the lakes studied. Comparisons between Chitu and nearby lakes formed obvious outliers.

Most Abundant Taxa and mRNA Transcripts

Figure 4 shows the distribution across habitats of abundant taxa at family rank or below, based on amplicon sequencing (prokaryotes and plastids only). It also lists RNA/DNA abundance ratios indicating the relative metabolic activity, number of OTUs and rRNA contigs. Together these cover 46–75% of total reads retrieved from each habitat. Table S3 lists the complete taxonomical composition for each rank and dataset. Few taxa were abundant in all six habitats, the deep-branching RF3 being an exception (min. abundance 0.8%). This phylum-level clade includes uncultured sequences from soda lakes, deep-sea sediments and enterosymbionts, with similarity to the prokaryotic genus incertae sedis Gemella [29]. Methanocalculus had the second highest average abundance, while Marine Group 1 Thaumarchaeota had the highest RNA/DNA-ratio and Rhodobacteraceae the highest diversity with 64 OTUs.

thumbnail
Figure 4. Distribution matrix with, DNA/RNA ratio, number of OTUs and rRNA contigs for the five most abundant taxa in each habitat.

Abundances are based on DNA amplicons from collection filters except those indicated with a star (*), instead based on prefilter-derived datasets. Taxa were defined at family level except for RF3 and MG I where information was not available at this resolution. DNA/RNA ratios are based on the dataset with highest RNA abundance and number of rRNA contigs include only those >750 bp. The dendogram indicate average linkage clustering of habitats based on OTU distribution (BC-dissimilarity).

https://doi.org/10.1371/journal.pone.0072577.g004

Table 2 lists all environmental datasets containing sequences most similar to amplicon or rRNA contigs from the abundant taxa included in Figure 4. This included datasets from seven alkaline lakes, eight saline or brackish-, and six non-saline bodies of water with unknown or neutral pH. It also included three datasets from soil and two cultured isolates: Rhodobaca bogoriensis [8] and a symbiont of Hydra magnipapillata [30]. Sequences derived from Mono Lake, California were the most commonly encountered amongst those most similar to abundant taxa.

thumbnail
Table 2. Sequences from environmental samples and cultured isolates similar to abundant taxonomic groups.

https://doi.org/10.1371/journal.pone.0072577.t002

Our complementary cDNA shotgun sequencing approach allowed insights into the most abundant eukaryotic taxa in lakes Beseka and Arenguadi (Table 3). The primers used also amplified chloroplast rRNA for all photosynthetic eukaryotes encountered, in some cases improving the insight into their distributions. For example, the most abundant eukaryotic family encountered in Arenguadi, Pavlovaceae, appeared more abundant at 2 m than at the surface of the lake and was also encountered in Chitu.

Further, diatoms from the family Thalassiosiraceae showing low abundance in the shotgun dataset were the most abundant eukaryotic taxon in the prefilter-derived dataset. Beseka appeared to harbor a contrasting eukaryotic community compared to Arenguadi, with phototrophs instead dominated by the mentioned diatoms, cryptophytes and Chlorophyceae. Non-phototrophic flagellates were present in both datasets, but with Placididea dominating in Beseka and Bicosoecidae in Arenguadi. Ciliates from different families were present in both lakes but more abundant in Arenguadi with Dysteriida constituting 2.8% of the sequences (see Table 3).

UniRef protein sequence clusters matching abundant putative mRNA-transcripts are listed in Table S4. Transcripts of Photosystems I and II were the most abundant of the genes with known functions (n = 39 in Arenguadi, n = 49 in Beseka). Various flagellar gene transcripts were also abundant in both lakes (n = {29, 12}). In addition, Arenguadi contained many transcripts from particulate methane monooxygenase (n = 26), others similar to a gene involved in calcium binding (A0L9Q4; n = 10), Chaperone DnaK (n = 7), and viral genes such as RNA-directed DNA polymerase (n = 10), RNA-directed RNA polymerase (n = 5), capsid and structural proteins (n = 4). Beseka instead contained transcripts from ammonia monooxygenase (amoA, n = 11) and nitrite reductase (nirK, n = 4).

Effect of Filtering

It is possible that the pre-filtering of water samples biased the community structure in collected biomass. Most OTUs encountered from collection-filters of the surface samples from Beseka were also encountered from the prefilters (Fig. S3), but often at contrasting abundances. In order to assess this “prefilter-bias”, taxon abundances were compared between the datasets derived from prefilters and corresponding collection-filters. A comparison was also made between the dataset derived from centrifugation (LAb C) and those from collection-filters. All consistent and significant outcomes of these comparisons are presented in Table 4. Eight of ten affected taxa showed decreased abundances, i.e. were more likely to pass through the prefilter. Seven of these were also less abundant in LAb C, at ratios indicating a more severe bias than from pre-filtering. Two taxa showed the opposite influence, i.e. increased abundance on prefilters, both from the phylum Planctomycetes.

thumbnail
Table 4. Families with relative abundance consistently influenced by filtering in lakes Arenguadi and Beseka.

https://doi.org/10.1371/journal.pone.0072577.t004

Several abundant taxa in prefilter-derived datasets were completely missing or uncommon in datasets from collection-filters. An example is Arthrospira platensis constituting 28% of the prefilter sequences in Chitu (Fig. 4, fam. Phormidiaceae; Table S3), indicating that most cells from these taxa could not pass the prefilters.

Discussion

Phylogenetic Diversity

All soda lakes studied harbored remarkably diverse microbial communities, considering their high pH. This also applied to pre-filter samples including filamentous or particle-associated organisms mostly missing from downstream collection filters. Surprisingly, the lake with the most extreme conditions (anoxic Lake Chitu) yielded the highest OTU richness, followed by the lake with the second highest pH and salinity (Abijata). The two anoxic samples from the stratified Lake Arenguadi also yielded relatively higher richness than surface samples. This is clearly a blatant transgression of the common notion that more extreme habitats should be less diverse. It even indicates the opposite: a positive correlation between phylogenetic richness and salinity or pH. To test this intriguing, counter-intuitive hypothesis properly, however, a larger number of replicates and lakes would be needed, evenly distributed along salinity and pH gradients.

In addition, cell density appeared correlated with diversity across depths in Arenguadi and when comparing to Lake Shalla, whose cell density was one order of magnitude lower (Figures 1, S2). Although our estimates were limited to these two lakes, a previous study estimated similar values of bacterial cells per volume in Arenguadi, placing Chitu and Abijata at about half its cell density, Shalla and Beseka about one order of magnitude below [31]. This agrees with the hypothesis that pH and salinity also increases richness. The effect these parameters have on productivity and prokaryotic cell density is challenging to disentangle and may involve complex trophic interactions, as grazers are generally more sensitive to salinity and pH. Although not measured in this study, the primary productivity rate is also expected to play an important role in these interactions.

The parametric richness estimation used compensates for contrasting sequencing depth, but relies on an assumption that sequence datasets constitute a representative subsample of the underlying biological community [28]. There are several problems with this assumption, including bias introduced from sampling, rRNA gene copy number [32], nucleic acid extraction [33] and PCR [34]. While these are expected to bias all amplicon datasets in a similar manner, cell density might not. However, concentrations of extracted nucleic acid did not follow the same trend as cell densities (Table S2), Further, template concentrations were adjusted prior to PCR, to avoid such bias. Thus, the correlation between diversity and cell density was likely not artificial.

The implicit richness definition used here was OTUs per volume unit, since the same sample volume was collected from each lake and mixed before filtering. Similar sample volumes were also filtered (Table S1). With larger cell density, we thus sampled more cells, more likely to represent higher richness. Rather than a sampling bias, this is arguably a general issue with comparing richness between habitats of contrasting biomass, area or volume [35].

Although no published studies utilised cloning-free high throughput sequencing to estimate the diversity of soda lake water samples, Xiong et al. [36] used it to analyze lake sediments, identifying a negative correlation between pH and richness. This disagrees with our hypothesis for planktonic diversity, but it is clearly possible that benthic communities show different correlations to these factors. Studies of Tibetan lakes [37] and the Baltic Sea [38] have examined similar salinity ranges as that studied here. Both identified a strong influence of salinity on community composition, but not richness. As opposed to pH and salinity, previous findings support the finding of anoxic hypolimnia being more diverse than corresponding epilimnia [17], [39], [40]. The cause of this is equally intriguing and possibly due to a high degree of endemism [41], challenging another common notion, namely that “everything is everywhere” [42].

OTU richness in surface [43] and hypolimnion [40] samples of neutral freshwater lakes has previously been studied using the same sequencing platform and noise-filtering as employed here (AmpliconNoise) [44]. The range of rarefied OTU richness from our soda lake datasets (Fig. S1) is approximately half of that obtained in these studies (300–600 at 5,000 reads in the former and 74–392 in the later). However, these studies targeted the V3–V4 regions of SSU rRNA instead of V5–V6, possibly resulting in higher richness estimates, not comparable to ours [45], [46]. It also appears that the pre-filtering used here prevented detection of several taxa. The observation of taxonomic richness comparable to neutral freshwater lakes agrees well with previous observations of morphological [3] and molecular diversity [20].

Richness of RNA-derived datasets was comparable to that of DNA-derived datasets in most samples, indicating that the majority of diversity originated from the active community, rather than an inactive “seed bank”. Although total active richness of RNA cannot theoretically be higher than that of available DNA, richness estimates suggested this in one of the samples (Arenguadi 2 m; Fig. 1). This may be explained by PCR bias affecting RNA-derived (cDNA) template in a different manner than the relatively longer DNA template. This was supported by RNA-derived datasets from Arenguadi having significantly higher rates of detected chimeras than DNA counterparts (Table S2). Artifacts introduced during reverse transcription may also have caused it.

Community Composition and Correlation to Physicochemical Parameters

Clustering and NMDS analyses supported both by OTU- and taxonomic composition, divide the datasets into six well-separated groups or habitats: one for each lake except Arenguadi, where epi- and hypolimnion were separated. Most of the dominant taxa show sharp abundance differences across habitats (Fig. 4) and shallow samples inside the same lake were significantly more similar to each other than those from different lakes. Likewise, relatively few OTUs were shared between more than one or two habitats (Fig. 2), compared to those shared between depths (Fig. S3) or spatial replicates. This difference was more pronounced in RNA-derived datasets. This is expected in an ecosystem where the activities of taxa (RNA) are more strongly influenced by local conditions than the DNA pool, also containing a “seed bank” of inactive and sporulating organisms and thus expected be more randomly distributed, spatially [47].

Out of OTUs shared between lakes (Fig. 2), two pairs: Abijata and Shalla; as well as Arenguadi and Chitu, showed larger overlaps between them than other lakes. The former overlap may be explained by the fact that Shalla and Abijata are located close together and were connected as recently as 2,000 years ago [48]. As for Arenguadi and Chitu, these were the only two lakes to contain samples from anoxic environments. Thus, obligate anaerobes were only shared between them out of the lakes studied. Except for Chitu, there is some support for the notion that sampled lakes located closer together harbored more similar communities than those far apart (Fig. S7). Although not significant, this could indicate distance-dependent dispersal limitations.

Out of the measured parameters, oxygen (presence or absence), pH, Na+ and K+ concentrations were significantly correlated to the OTU composition in the studied habitats. Although oxygen appeared to have the largest influence when partitioning the compositional variation in pooled habitat-datasets, Na+ was equally important when including individual datasets (Fig. S6). The relative contributions of pH and K+ were equally hard to disentangle. Regardless of model used, these parameters explain about 30% of community variation. As mentioned, salinity and oxygen have previously been established as important factors for shaping the microbial composition in aquatic habitats [16], [38], [41]. Na+ and pH have also been indicated as important influences for OTU composition in soda lake sediments [36].

Taxa Encountered and Possible Ecological Roles

We expect that the amplicon datasets obtained were representative for the majority of taxa in the underlying community of bacterial and archaeal plankton. This was confirmed using complementary shotgun sequencing, alternative DNA extraction- and harvesting protocols, for the pre-filter- and LAb C samples. Resulting datasets conformed to habitat-specific clustering patterns and shared similar abundances for most taxa, compared to corresponding amplicon datasets from default protocols. Exceptions include Arthrospira, Thalassiosiraceae and Planctomycetes, whose abundances were dramatically decreased by pre-filtering. This is expected, considering these taxa have filamentous growth, large rigid cells and attach to surfaces or other cells, respectively. Correspondingly, underrepresented taxa are good candidates for having smaller than average cells and it appears these were not successfully collected using centrifugation.

Due to filtering bias, it was challenging to identify the main primary producers in the lakes studied. Arthrospira platensis appeared to dominate the surface of Chitu. This was also the only lake with large flocks of Lesser Flamingos present during sampling. These birds are typically found together with Arthrospira, which is their main diet [49]. This genus was only present in trace amounts in Arenguadi, consistent with earlier reports that it is disappearing from the lake [50]. Instead, abundance of photosynthetic taxa was dominated by the Cyanobacterial genera Leptolyngbya and Anabaenopsis, but mainly by the eukaryotic haptophyte Pavlovaceae. The later family is a flagellated unicellular algae commonly found in brackish littoral costal waters [51]. Chloroplastic 16S from Pavlovaceae was also present in other lakes, but two orders of magnitude less abundant. At genus rank, most reads of this family were classified as Pavlova, while the only full-length 18S rRNA contig obtained from the taxon was more similar (99%) to Diacronema (AF106056). However, these two genera appear polyphyletic and a merger has been suggested [51].

No cyanobacteria were detected in Abijata, Shalla or Beseka, probably due to filtering bias. Given the lack of cyanobacterial reads, it is probable that photosynthesis in Beseka was dominated by eukaryotes. Compared to Arenguadi, a different and more diverse community of photosynthetic eukaryotes was present, dominated by cryptophytes in the families Geminigeraceae and Chroomonadaceae. Mostly studied as model organisms for secondary endosymbiosis, these are flagellated and unicellular, like the Pavlovaceae. Also abundant were Chlorophyceae, mainly unclassified at higher ranks, and diatoms of the family Thalassiosiraceae.

Anoxygenic photosynthesis also appeared to contribute to primary production in several lakes. A diversity of non-sulfur purple bacteria from the family Rhodobaceraceae (genera Rhodobaca, Rhodobacter, Pseudorhodobacter and Roseibacter) dominated in Abijata and Shalla, while purple sulfur bacteria from the genus Ectothiorhodospira dominated in the anoxic lake Chitu and also occurred in Abijata. The non-phototrophic genus Thioalkalivibrio in the same family (Ectothiorhodospiraceae) was abundant in lakes Chitu and Arenguadi. An internal sulfur cycle is suggested by the presence of sulfate reducers from the families Desulfohalobiaceae (mainly Desulfonatronovibrio), as previously observed in soda lakes [3]. Both Thioalkalivibrio and Desulfonatronovibrio are known as widespread and diverse groups commonly found in soda lakes [11]. Most similar environmental sequences from other studies were also from soda lakes. No obvious sulfate reducers could be identified in Arenguadi. It is possible that hydrothermal springs feeding some of the lakes studied contain sulfide of geological origin, although no studies supporting this could be identified.

Methanogens, mainly from the genera Methanocalculus Methanolobus and Methanoseata were found, with the highest relative abundance in Arenguadi (at 30 m) and Chitu. A single OTU classified as Methanocalculus dominated among these, most similar to isolates from a soda lake on the Kulunda Steppe (Table 2). The most similar (98%) validly described isolate was M. halotolerans, a hydrogenotrophic and methylotrophic species isolated from an oilfield [52]. Aerobic methane oxidation in the surface of Arenguadi was evident from mRNA transcripts and presence of Methylomicrobium, previously encountered in soda lakes [13].

Ammonia-oxidizing archaea (AOA) from Marine Group I Thaumarchaeota [53] were found in high abundance in lakes Shalla and Beseka, constituting as much as half of the RNA-derived reads at 13 m depth in Beseka while also abundant at the surface. Although sequences from AOA inhabiting soda lakes exist [54], those encountered here were more similar to environmental sequences from two different gold mines [55], [56] (Table 2), belonging to the terrestrial subgroup Lambda I [57]. Active ammonia oxidation was confirmed by active transcription of archaeal amoA, as well as nirK. The later observation is particularly interesting as it supports the suggested role of AOA in ‘nitrifier denitrification’ [58], recently demonstrated in soil [59], marine habitats [60] and enrichment cultures [61] including estuaries with similar salinity to Lake Beseka [62].

It is possible that Planctomycetes were involved in anaerobic ammonia oxidation (anammox), although none of the known anammox taxa [63] were encountered. The type species of the most common order found (Phycisphaerales) is instead a heterotrophic algae symbiont [64]. It is possible that nitrogen fixation is carried out by Rhodobacter in Lake Shalla, Azoarcus (fam. Rhodocyclaceae) in Arenguadi and Derxia in Beseka (as well as other taxa in fam. Alcaligenaceae). Putative denitrifiers include Rhodobacteraceae.

Other taxa encountered at high abundance include aerobic heterotrophs (e.g. Bacteroidetes, Moraxellaceae, Marinicella) and fermentative anaerobes (e.g. Thermoplasmatales). Taxa typical for highly specialized metabolisms were also encountered such as Oceanospirillaceae and Nitriliiruptor, the later known for being able to catabolize nitriles or cyanides [65]. Others, like RF3, remain poorly studied and with unknown function. Many in both categories showed high similarity to sequences found previously in saline or soda lakes (Table 2).

A diversity of putatively bacterivorous eukaryotes was present including ciliates (Dysteriida, Cyclidiidae, Didiniidae etc.), flagellates (Bicosoecidae, Placididea, Colpodella etc.), rotifers (Polyarthra, Brachionus), Simocephalus, Cercozoa and Heterolobosea, the most abundant listed in Table 3. Considering their abundance and diversity, it is probable that these exert a considerable top-down control on the prokaryotic community. To what extent viruses control the diversity and structure of the microbial community can only be guessed. A few putative bacteriophage transcripts were found among the limited mRNA reads from Lake Arenguadi. Transcripts from information processing genes were found in higher abundance, however, from (+)ssRNA- and retrovirus, groups known to only infect eukaryotes.

Materials and Methods

Sample Collection, Filtering and Storage

Sampling in Abijata-Shalla National Park was carried out with permission and supervision from the Ethiopian Wildlife Conservation Authority. No permission was required for the other two lakes (Beseka and Arenguadi), located in publicly accessible areas.

Water samples (excluding Abijata sample C; “LAb C”) were collected in March 2011 using a 2.5L Niskin bottle (Ocean Scientific International Ltd.), kept in sealed containers and pre-filtered using 5 µm polycarbonate filters (Poretics Ø47 mm, Osmonics Inc. USA.) in order to avoid immediate clogging of more narrow collection filters suitable for prokaryotic cells. The filtrate was then passed through 0.2 µm Sterivex™ columns (Millipore) until clogging occurred, in order to maximize cell yield. Site names, coordinates, depths and filtered volumes are listed in Table S1. While prefilters were deposited in 15 mL Falcon tubes filled with RNALater, Sterivex columns were filled with RNALater and sealed. All samples were stored at 4°C until further processing.

Sample LAb C was collected in December 2011 and processed using a different, more rapid protocol, mainly for evaluation purposes. Surface water was collected using sterile 50 ml Falcon tubes, transported on ice to Addis Ababa University, then preserved at 4°C for less than a week. Isohaline PBS (pH 11) was added to the sample and biomass harvested by centrifugation at 3700 RPM for 30 min at 4°C from 200 ml water by repeatedly removing supernatant and adding new water using a Consul 21R centrifuge (Orto Alresa). Finally, cell pellets were washed with PBS and centrifuged twice to remove salt particles. Pure cell pellets were preserved at −20°C until DNA extraction.

Measurements of Physicochemical Parameters

Concentrations of Na+ and K+ and a number of other ions were measured from native surface water samples (stored at 4°C in 15 mL Falcon tubes), using inductively coupled plasma optical emission spectrometry (Elemental IRIS, Thermo Fisher Scientific Inc.). Salinity, pH and dissolved oxygen (DO) was measured on site during sampling: total salinity with a standard refractometer (0–100‰, ATAGO Co. Ltd.); pH with a portable pH-meter (Oakton pH 110, Eutech Instruments Pty. Ltd.) and confirmed with indicator strips (Merck, range 5–10); and DO using a portable dissolved oxygen meter (Hi9143, Hanna Instruments). Due to equipment failure, the oxygen level could not be measured properly in lakes Abijata, Beseka or Shalla. To compensate, DO was instead treated as a binary variable in future analysis (presence or absence), based on read-outs and earlier measurements. All sampled depths in the problematic lakes were determined as nearly saturated (presence).

Cell Enumeration

Unfiltered water samples were collected in 15 ml Falcon tubes and filled with formaldehyde to a concentration of 2%. DAPI staining was used for enumeration of total prokaryotic cells. Formaldehyde-fixed water samples were thoroughly mixed by vortexing, 1 ml aliquots dried on 0.2 µm filters, incubated with 2% DAPI solution for 15 min in the dark, then rinsed with sterile distilled water (2×10 ml). Filters mounted on microscope slides were inspected using a Zeiss Axioplan fluorescent microscope and manually counted in diagonal squares of an overlaid grid. Mean and standard errors of cell densities were calculated using a minimum of 12 squares per sample.

Nucleic Acid Extractions

DNA and RNA was simultaneously extracted directly from Sterivex columns using the AllPrep DNA/RNA Extraction Kit (Qiagen). Prior to extraction, columns were opened, RNALater removed and replaced with lysis buffer (RLT Plus). The columns were then re-sealed, rotated gently and incubated for 1 minute before lysate was passed through the filter by manual air pressure application using a syringe. Subsequent extraction steps were carried out according to the manufacturer’s protocol and extracts stored at −80°C. From LAb C and prefilters, community DNA was extracted using CTAB as described previously [66]. Extracted DNA pellets were dried and resuspended in 50 µl of TE buffer (pH 8) and stored at −20°C. Nucleic acid concentrations were determined using NanoDrop™ spectrometry.

cDNA Synthesis

Total RNA was quality assessed using gel electrophoresis. Extracts where RNA was detected, while lacking well-contrasted bands corresponding to the two ribosomal subunits were discarded, retaining only those from Arenguadi and Beseka. From these lakes, single-strand reverse transcription was carried out to provide template for amplicon libraries. Superscript III (Invitrogen) was used according to the manufacturer’s protocol, random hexamer primed and with subsequent RNAse H digestion. In addition, the two surface samples were subjected to double-stranded cDNA synthesis as described previously [23].

Amplicon Library Preparation

PCR amplification of the V5–V8 region of prokaryotic SSU rRNA (16S) was carried out from extracted DNA and single-stranded cDNA using the primers Uni787F (5′-ATTAGATACCCNGGTAG-3′) and Uni1492R (50-GNTACCTTGTTACGACTT-30) [67] using a two-step (nested) PCR protocol described previously [68]. Template concentrations and number of PCR cycles (Table S2) were adjusted to achieve equal concentrations of final products. Triplicate PCR reactions were pooled and purified using GenElute PCR Clean-Up kit (Sigma) prior to the second PCR step, instead using primers with attached sample-specific, error-correcting barcodes (“multiplex identifiers”) and GS-FLX adaptors (Lib-L). Resulting amplicons were cleaned using AMPure XP (Beckman Coulter) following the manufacturer’s protocol (bead-to-sample ratio 9∶10). Amplicon DNA was analyzed using gel electrophoresis to ensure complete removal of primers and negligible amounts of non-barcoded product. Concentrations were measured using Qubit and amplicons stored at −80°C until pooling in equimolar amounts and sequencing.

Sequencing and Data Submission

Pyrosequencing, ds-cDNA synthesis and shotgun library preparation was carried out at the Norwegian High-Throughput Sequencing Centre. Amplicons were sequenced using GS-FLX Titanium chemistry (Lib-L) and cDNA shotgun libraries using GS-FLX+. No fragmentation was carried out since sequences longer than 3000 bp (DNA-contamination) were rare. Resulting flowgrams were submitted to the NCBI Sequence Read Archive with study accession number SRA061754.

Sequencing Processing, Including Filtering and Noise-removal

In amplicon datasets, filtering, removal of noise and chimeric sequences was carried out using AmpliconNoise (AN) [44]. This method shows the most complete removal of PCR and sequencing artifacts, while not obfuscating real, OTUs [69]. Barcode and primer sequences were removed and resulting sequences annotated with read-abundance. In addition to the chimera filtering carried in AN (Perseus), UCHIME [70] was used to remove any remaining chimeric sequences (min. score 0.1) and SilvaMod106 as reference database [29]. The resulting “cleaned” sequences were clustered into OTUs using maximum linkage based on pairwise Needleman-Wunsch alignment distances at a 3% distance cutoff using AN [44]. Diversity indices (1-D and H’) were calculated from resulting OTUs using the OTUDist.sh script distributed with AN (v1.26 alpha) and rarefaction carried out using the program E-Rarefaction [28]. Rarefied richness was based on the smallest dataset, excluding the Chitu prefilter (2,967 reads). Shotgun cDNA reads were filtered by removing reads shorter than 150 bp, with degenerate bases (‘Ns’) or average quality below 25.

Cleaned amplicon sequences and filtered shotgun reads were subjected to taxonomic classification using CREST [29]. Assembly of full-length rRNA contigs was carried out independently using shotgun reads from taxonomic groups as described previously [71]. Shotgun reads with an alignment bitscore below 50 were screened and cleaned for ncRNAs using Infernal and Rfam [72], [73], then aligned to UniRef90 [74] using BLASTX to identify putative mRNA transcripts (min. bitscore 45).

Ordination, Variation Partitioning and Other Statistical Analyses

Calculation of Bray-Curtis dissimilarities between datasets as well as hierarchical clustering, NMDS, parameter correlation and variation partitioning based on these, were carried out using the R programming language [75] and the Vegan package [76]. To support OTU-based analysis, taxonomic groups were derived from the number of reads assigned to each taxon at all ranks from domain to genus using the composite All_Composition.txt output from CREST [29]. Taxonomic comparison of datasets derived from prefilters vs. collection filters, as well as shotgun sequencing vs. amplicons, was carried out as described previously [22].

Supporting Information

Figure S1.

Rarefaction curves of OTUs from amplicon samples. The number of encountered OTUs (perceived richness) is plotted relative to sub-sampled sequence datasets size, i.e. number of reads. For Chitu and Abijata, samples of same depth pooled in silico are plotted in addition to individual ones.

https://doi.org/10.1371/journal.pone.0072577.s001

(TIF)

Figure S2.

Cell density at different depths of Lake Arenguadi estimated using DAPI staining. Grey squares indicate averages between counts and the solid lines 95% confidence intervals.

https://doi.org/10.1371/journal.pone.0072577.s002

(EPS)

Figure S3.

Venn diagrams showing the distributions of shared OTUs within lakes. Diagrams are annotated with the number of OTUs shared for each possible subset within (A) all amplicon datasets from a depths of 0 m in Lake Beseka; (B) all DNA amplicon datasets (excluding prefilter sample) and (C) all cDNA amplicon datasets across depths in Lake Beseka; (D) all DNA- and (E) all cDNA amplicon datasets in Lake Arenguadi between 0 to 3 meters.

https://doi.org/10.1371/journal.pone.0072577.s003

(EPS)

Figure S4.

Average linkage clustering dendogram of amplicon sequence datasets using Bray-Curtis dissimilarity.

https://doi.org/10.1371/journal.pone.0072577.s004

(EPS)

Figure S5.

NMDS based on Bray Curtis dissimilarities between relative abundances of taxonomic groups.

https://doi.org/10.1371/journal.pone.0072577.s005

(EPS)

Figure S6.

Venn-diagrams illustrating partitioning of variation of selected physicochemical variables. Partitioning of total community. As response variable, OTU abundance across individual sequence datasets was used in (A) and OTU abundance across habitats (pooled datasets) in (B).

https://doi.org/10.1371/journal.pone.0072577.s006

(EPS)

Figure S7.

Bray-Curtis community dissimilarity between surface samples from different lakes plotted vs. the physical distance between lakes. Where replicate surface samples existed, the average composition was used. Minimum distances between lakes were measured using Google Maps.

https://doi.org/10.1371/journal.pone.0072577.s007

(PDF)

Table S1.

Full list of sampling sites and environmental parameters measured during sampling.

https://doi.org/10.1371/journal.pone.0072577.s008

(DOCX)

Table S3.

Taxonomic composition of all sequence datasets.

https://doi.org/10.1371/journal.pone.0072577.s010

(XLS)

Table S4.

UniRef90 protein clusters. Accession, name and length of representative protein sequence for UniRef with number of putative mRNA reads sharing best alignments to each cluster, mean read length and coverage (total read length/protein sequence length in bp). Only clusters >2 mRNA reads and aligning with a bitscore >45 included.

https://doi.org/10.1371/journal.pone.0072577.s011

(XLSX)

Acknowledgments

We would like to thank Baye Sitotaw, Ingrid Mørkeseth, Yemisirach Mulugeta and all of the rest of the team from Addis Ababa University for sampling, driving and planning during sample collection, in particular for an unexpected and brave dive head-first into the murky waters of Abijata. Thanks to Ingunn H. Thorseth for performing the ICP analyses, and also to Mia Bengtsson and Hallgjerd Eydal for expertise and discussions. Ave Tooming Klunderud and others at the Norwegian Sequencing Centre are acknowledged for invaluable help with sequencing and library preparation.

Author Contributions

Conceived and designed the experiments: AL LØ AS AG. Performed the experiments: AL AS DC. Analyzed the data: AL AS IJ LØ. Wrote the paper: AL.

References

  1. 1. Grant WD (2006) Alkaline environments and biodiversity. In: Gerday EC, Glansdorff N, editors. Extremophiles. Oxford, UK: UNESCO, Eolss Publishers.
  2. 2. Melack JM, Kilham P (1974) Photosynthetic rates of phytoplankton in East African alkaline, saline lakes. Limnol Oceanogr 19: 743–755.
  3. 3. Zavarzin GA, Zhilina TN (2000) Anaerobic chemotrophic alkaliphiles. In: Seckbach J, editor. Journey to Diverse Microbial Worlds: Adaptation to Exotic Environments: Kluwer Academic Publishers. 191–108.
  4. 4. Oren A (1999) Bioenergetic aspects of halophilism. Microbiol Mol Biol Rev 63: 334–348.
  5. 5. Zhilina TN, Zavarzina DG, Panteleeva AN, Osipov GA, Kostrikina NA, et al. (2012) Fuchsiella alkaliacetigena gen. nov., sp. nov., an alkaliphilic, lithoautotrophic homoacetogen from a soda lake. Int J Syst Evol Microbiol 62: 1666–1673.
  6. 6. Ballot A, Krienitz L, Kotut K, Wiegand C, Metcalf JS, et al. (2004) Cyanobacteria and cyanobacterial toxins in three alkaline Rift Valley lakes of Kenya - Lakes Bogoria, Nakuru and Elmenteita. J Plankton Res 26: 925–935.
  7. 7. Asao M, Pinkart HC, Madigan MT (2011) Diversity of extremophilic purple phototrophic bacteria in Soap Lake, a Central Washington (USA) Soda Lake. Environ Microbiol 26: 925–935.
  8. 8. Milford AD, Achenbach LA, Jung DO, Madigan MT (2000) Rhodobaca bogoriensis gen. nov. and sp. nov., an alkaliphilic purple nonsulfur bacterium from African Rift Valley soda lakes. Arch Microbiol 174: 18–27.
  9. 9. Medová H, Boldareva EN, Hrouzek P, Borzenko SV, Namsaraev ZB, et al. (2011) High abundances of aerobic anoxygenic phototrophs in saline steppe lakes. FEMS Microbiol Ecol 76: 393–400.
  10. 10. Rees HC, Grant WD, Jones BE, Heaphy S (2004) Diversity of Kenyan soda lake alkaliphiles assessed by molecular methods. Extremophiles 8: 63–71.
  11. 11. Sorokin DY, Kuenen JG, Muyzer G (2011) The microbial sulfur cycle at extremely haloalkaline conditions of soda lakes. Front Microbiol 2: 44.
  12. 12. Antony CP, Murrell JC, Shouche YS (2012) Molecular diversity of methanogens and identification of Methanolobus sp. as active methylotrophic Archaea in Lonar crater lake sediments. FEMS Microbiol Ecol 81: 43–51.
  13. 13. Surakasi VP, Antony CP, Sharma S, Patole MS, Shouche YS (2010) Temporal bacterial diversity and detection of putative methanotrophs in surface mats of Lonar crater lake. J Basic Microbiol 50: 465–474.
  14. 14. Antony CP, Doronina NV, Boden R, Trotsenko YA, Shouche YS, et al. (2012) Methylophaga lonarensis sp. nov., a moderately haloalkaliphilic methylotroph isolated from the soda lake sediments of a meteorite impact crater. Int J Syst Evol Microbiol 62: 1613–1618.
  15. 15. Lin J-L, Radajewski S, Eshinimaev BT, Trotsenko YA, McDonald IR, et al. (2004) Molecular diversity of methanotrophs in Transbaikal soda lake sediments and identification of potentially active populations by stable isotope probing. Environ Microbiol 6: 1049–1060.
  16. 16. Wu Q, Chatzinotas A, Wang J, Boenigk J (2009) Genetic Diversity of Eukaryotic Plankton Assemblages in Eastern Tibetan Lakes Differing by their Salinity and Altitude. Microb Ecol 58: 569–581.
  17. 17. Humayoun SB, Bano N, Hollibaugh JT (2003) Depth distribution of microbial diversity in Mono Lake, a meromictic soda lake in California. Appl Environ Microbiol 69: 1030–1042.
  18. 18. Dimitriu PA, Pinkart HC, Peyton BM, Mormile MR (2008) Spatial and temporal patterns in the microbial diversity of a meromictic soda lake in Washington State. Appl Environ Microbiol 74: 4877–4888.
  19. 19. Wang J, Yang D, Zhang Y, Shen J, van der Gast C, et al. (2011) Do patterns of bacterial diversity along salinity gradients differ from those observed for macroorganisms? PLoS ONE 6: e27597.
  20. 20. Mesbah N, Abou-El-Ela S, Wiegel J (2007) Novel and Unexpected Prokaryotic Diversity in Water and Sediments of the Alkaline, Hypersaline Lakes of the Wadi An Natrun, Egypt. Microbial Ecology 54: 598–617.
  21. 21. King AJ, Freeman KR, McCormick KF, Lynch RC, Lozupone C, et al. (2010) Biogeography and habitat modelling of high-alpine bacteria. Nat Commun 1: 53.
  22. 22. Lanzén A, Jørgensen SL, Bengtsson MM, Jonassen I, Øvreås L, et al. (2011) Exploring the composition and diversity of microbial communities at the Jan Mayen hydrothermal vent field using RNA and DNA. FEMS Microbiol Ecol 77: 577–589.
  23. 23. Urich T, Lanzén A, Qi J, Huson DH, Schleper C, et al. (2008) Simultaneous assessment of soil microbial community structure and function through analysis of the meta-transcriptome. PLoS ONE 3: e2527.
  24. 24. Baxter RM (2002) Lake Morphology and Chemistry. In: Tudorancea C, Taylor WD, editors. Ethiopian rift valley lakes: Backhuys.
  25. 25. Ayenew T (2007) Water management problems in the Ethiopian rift: Challenges for development. J Afr Earth Sci 48: 222–236.
  26. 26. Gloaguen AG, Egbert J (2009) Richard (2009) Non-climatic growth of the saline Lake Beseka, Main Ethiopian Rift. J Arid Environ 73: 287–295.
  27. 27. Gebre-Mariam Z (1998) Human Interactions and Water Quality in the Horn of Africa. In: Schoneboom J, editor. AAAS, Washington, DC.
  28. 28. Quince C, Curtis TP, Sloan WT (2008) The rational exploration of microbial diversity. ISME J 2: 997–1006.
  29. 29. Lanzén A, Jørgensen SL, Huson DH, Gorfer M, Grindhaug SH, et al. (2012) CREST - Classification Resources for Environmental Sequence Tags. PLoS ONE 7: e49334.
  30. 30. Chapman JA, Kirkness EF, Simakov O, Hampson SE, Mitros T, et al. (2010) The dynamic genome of Hydra. Nature 464: 592–596.
  31. 31. Zinabu GM, Taylor WD (1997) Bacteria-chlorophyll relationships in Ethiopian lakes of varying salinity: are soda lakes different? J Plankton Res 19: 647–654.
  32. 32. Kembel SW, Wu M, Eisen JA, Green JL (2012) Incorporating 16S Gene Copy Number Information Improves Estimates of Microbial Diversity and Abundance. PLoS Comput Biol 8: e1002743.
  33. 33. Terrat S, Christen R, Dequiedt S, Lelièvre M, Nowak V, et al. (2012) Molecular biomass and MetaTaxogenomic assessment of soil microbial communities as influenced by soil DNA extraction procedure. Microb Biotechnol 5: 135–141.
  34. 34. Suzuki MT, Giovannoni SJ (1996) Bias caused by template annealing in the amplification of mixtures of 16S rRNA genes by PCR. Appl Environ Microbiol 62: 625–630.
  35. 35. Gotelli NJ, Colwell RK (2001) Quantifying biodiversity: procedures and pitfalls in the measurement and comparison of species richness. Ecol Lett 4: 379–391.
  36. 36. Xiong J, Liu Y, Lin X, Zhang H, Zeng J, et al. (2012) Geographic distance and pH drive bacterial distribution in alkaline lake sediments across Tibetan Plateau. Environ Microbiol 14: 2457–2466.
  37. 37. Wu QL, Zwart G, Schauer M, van Agterveld MPK, Hahn MW (2006) Bacterioplankton community composition along a salinity gradient of sixteen high-mountain lakes located on the Tibetan Plateau, China. Appl Environ Microbiol 72: 5478–5485.
  38. 38. Herlemann DP, Labrenz M, Jürgens K, Bertilsson S, Waniek JJ, et al. (2011) Transitions in bacterial communities along the 2000 km salinity gradient of the Baltic Sea. ISME J 5: 1571–1579.
  39. 39. Klepac-Ceraj V, Hayes CA, Gilhooly WP, Lyons TW, Kolter R, et al. (2012) Microbial diversity under extreme euxinia: Mahoney Lake, Canada. Geobiology 10: 223–235.
  40. 40. Peura S, Eiler A, Bertilsson S, Nykänen H, Tiirola M, et al. (2012) Distinct and diverse anaerobic bacterial communities in boreal lakes dominated by candidate division OD1. ISME J 6: 1640–1652.
  41. 41. Barberán A, Casamayor EO (2011) Euxinic freshwater hypolimnia promote bacterial endemicity in continental areas. Microb Ecol 61: 465–472.
  42. 42. O’Malley MA (2008) ‘Everything is everywhere: but the environment selects’: ubiquitous distribution and ecological determinism in microbial biogeography. Stud Hist Philos Biol Biomed Sci 39: 314–325.
  43. 43. Logue JB, Langenheder S, Andersson AF, Bertilsson S, Drakare S, et al. (2012) Freshwater bacterioplankton richness in oligotrophic lakes depends on nutrient availability rather than on species-area relationships. ISME J 6: 1127–1136.
  44. 44. Quince C, Lanzén A, Davenport RJ, Turnbaugh PJ (2011) Removing noise from pyrosequenced amplicons. BMC BIoinformatics 12: 38.
  45. 45. Jeraldo P, Chia N, Goldenfeld N (2011) On the suitability of short reads of 16S rRNA for phylogeny-based analyses in environmental surveys. Environ Microbiol 13: 3000–3009.
  46. 46. Youssef N, Sheik CS, Krumholz LR, Najar FZ, Roe BA, et al. (2009) Comparison of species richness estimates obtained using nearly complete fragments and simulated pyrosequencing-generated fragments in 16S rRNA gene-based environmental surveys. Appl Environ Microbiol 75: 5227–5236.
  47. 47. Jones SE, Lennon JT (2010) Dormancy contributes to the maintenance of microbial diversity. Proc Natl Acad Sci U S A 107: 5881–5886.
  48. 48. Benvenuti M, Carnicelli S, Belluomini G, Dainelli N, Di Grazia S, et al. (2002) The Ziway-Shala lake basin (main Ethiopian rift, Ethiopia): a revision of basin evolution with special reference to the Late Quaternary. J Afr Earth Sci 35: 247–269.
  49. 49. Owino AO, Oyugi JO, Nasirwa OO, Bennun LA (2001) Patterns of variation in waterbird numbers on four Rift Valley lakes in Kenya, 1991–1999. Hydrobiologia 458: 45–53.
  50. 50. Girma MB, Kifle D, Jebessa H (2012) Deep underwater seismic explosion experiments and their possible ecological impact - The case of Lake Arenguade - Central Ethiopian highlands. Limnologica (Other) 42: 212–219.
  51. 51. Bendif EM, Probert I, Hervé A, Billard C, Goux D, et al. (2011) Integrative taxonomy of the Pavlovophyceae (Haptophyta): a reassessment. Protist 162: 738–761.
  52. 52. Ollivier B, Fardeau ML, Cayol JL, Magot M, Patel BK, et al. (1998) Methanocalculus halotolerans gen. nov., sp. nov., isolated from an oil-producing well. Int J Syst Bacteriol 48 Pt 3: 821–828.
  53. 53. Pester M, Schleper C, Wagner M (2011) The Thaumarchaeota: an emerging view of their phylogeny and ecophysiology. Curr Opin Microbiol 14: 300–306.
  54. 54. Carini SA, Joye SB (2008) Nitrification in Mono Lake, California: Activity and community composition during contrasting hydrological regimes. Limnol Oceanogr 53: 2546–2557.
  55. 55. Rastogi G, Stetler L, Peyton B, Sani R (2009) Molecular analysis of prokaryotic diversity in the deep subsurface of the former Homestake gold mine, South Dakota, USA. J Microbiol 47: 371–384.
  56. 56. Nunoura T, Hirayama H, Takami H, Oida H, Nishi S, et al. (2005) Genetic and functional properties of uncultivated thermophilic crenarchaeotes from a subsurface gold mine as revealed by analysis of genome fragments. Environ Microbiol 7: 1967–1984.
  57. 57. Jørgensen SL, Hannisdal B, Lanzén A, Baumberger T, Flesland K, et al. (2012) Correlating microbial community profiles with geochemical data in highly stratified sediments from the Arctic Mid-Ocean Ridge. Proc Natl Acad Sci U S A 109: e2846–e2855.
  58. 58. Francis CA, Roberts KJ, Beman JM, Santoro AE, Oakley BB (2005) Ubiquity and diversity of ammonia-oxidizing archaea in water columns and sediments of the ocean. Proc Natl Acad Sci U S A 102: 14683–14688.
  59. 59. Bartossek R, Nicol GW, Lanzen A, Klenk H-P, Schleper C (2010) Homologues of nitrite reductases in ammonia-oxidizing archaea: diversity and genomic context. Environ Microbiol 12: 1075–1088.
  60. 60. Löscher CR, Kock A, Könneke M, LaRoche J, Bange HW, et al. (2012) Production of oceanic nitrous oxide by ammonia-oxidizing archaea. Biogeosciences 9: 2419–2429.
  61. 61. Santoro AE, Buchwald C, McIlvin MR, Casciotti KL (2011) Isotopic signature of N(2)O produced by marine ammonia-oxidizing archaea. Science 333: 1282–1285.
  62. 62. Mosier AC, Lund MB, Francis CA (2012) Ecophysiology of an ammonia-oxidizing archaeon adapted to low-salinity habitats. Microb Ecol 64: 955–963.
  63. 63. Fuchsman CA, Staley JT, Oakley BB, Kirkpatrick JB, Murray JW (2012) Free-living and aggregate-associated Planctomycetes in the Black Sea. FEMS Microbiol Ecol 80: 402–416.
  64. 64. Fukunaga Y, Kurahashi M, Sakiyama Y, Ohuchi M, Yokota A, et al. (2009) Phycisphaera mikurensis gen. nov., sp. nov., isolated from a marine alga, and proposal of Phycisphaeraceae fam. nov., Phycisphaerales ord. nov. and Phycisphaerae classis nov. in the phylum Planctomycetes. J Gen Appl Microbiol 55: 267–275.
  65. 65. Sorokin DY, van Pelt S, Tourova TP, Evtushenko LI (2009) Nitriliruptor alkaliphilus gen. nov., sp. nov., a deep-lineage haloalkaliphilic actinobacterium from soda lakes capable of growth on aliphatic nitriles, and proposal of Nitriliruptoraceae fam. nov. and Nitriliruptorales ord. nov. Int J Syst Evol Microbiol 59: 248–253.
  66. 66. Bengtsson MM, Sjøtun K, Øvreås L (2010) Seasonal dynamics of bacterial biofilms on the kelp Laminaria hyperborea. Aquatic Microbial Ecology 60: 71–83.
  67. 67. Roesch LFW, Fulthorpe RR, Riva A, Casella G, Hadwin AKM, et al. (2007) Pyrosequencing enumerates and contrasts soil microbial diversity. ISME J 1: 283–290.
  68. 68. Bengtsson MM, Sjotun K, Lanzen A, Ovreas L (2012) Bacterial diversity in relation to secondary production and succession on surfaces of the kelp Laminaria hyperborea. ISME J.
  69. 69. Lee CK, Herbold CW, Polson SW, Wommack KE, Williamson SJ, et al. (2012) Groundtruthing Next-Gen Sequencing for Microbial Ecology - Biases and Errors in Community Structure Estimates from PCR Amplicon Pyrosequencing. PLoS ONE 7: e44224.
  70. 70. Edgar RC, Haas BJ, Clemente JC, Quince C, Knight R (2011) UCHIME improves sensitivity and speed of chimera detection. Bioinformatics 27: 2194–2200.
  71. 71. Radax R, Rattei T, Lanzen A, Bayer C, Rapp HT, et al. (2012) Metatranscriptomics of the marine sponge Geodia barretti: tackling phylogeny and function of its microbial community. Environmental Microbiology 14: 1308–1324.
  72. 72. Gardner PP, Daub J, Tate J, Moore BL, Osuch IH, et al. (2011) Rfam: Wikipedia, clans and the “decimal” release. Nucleic Acids Res 39: D141–D145.
  73. 73. Nawrocki EP, Kolbe DL, Eddy SR (2009) Infernal 1.0: inference of RNA alignments. Bioinformatics 25: 1335–1337.
  74. 74. Suzek BE, Huang H, McGarvey P, Mazumder R, Wu CH (2007) UniRef: comprehensive and non-redundant UniProt reference clusters. Bioinformatics 23: 1282–1288.
  75. 75. R Development Core Team (2012) R: A Language and Environment for Statistical Computing. Vienna, Austria: R Foundation for Statistical Computing.
  76. 76. Oksanen J, Blanchet F, Kindt R, Legendre P, Minchin P, et al.. (2012) vegan: Community Ecology Package.
  77. 77. Wani AA, Surakasi VP, Siddharth J, Raghavan RG, Patole MS, et al. (2006) Molecular analyses of microbial diversity associated with the Lonar soda lake in India: an impact crater in a basalt area. Res Microbiol 157: 928–937.
  78. 78. Sorokin DY, Rusanov II, Pimenov NV, Tourova TP, Abbas B, et al. (2010) Sulfidogenesis under extremely haloalkaline conditions in soda lakes of Kulunda Steppe (Altai, Russia). FEMS Microbiol Ecol 73: 278–290.
  79. 79. Glatz RE, Lepp PW, Ward BB, Francis CA (2006) Planktonic microbial community composition across steep physical/chemical gradients in permanently ice-covered Lake Bonney, Antarctica. Geobiology 4: 53–67.
  80. 80. Zhang X, Kong F (2010) [Bacterial diversity in Zabuye salt lake of tibet by culture-independent approaches]. Wei Sheng Wu Xue Bao 50: 334–341.
  81. 81. Dillon J, McMath L, Trout A (2009) Seasonal changes in bacterial diversity in the Salton Sea. Hydrobiologia 632: 49–64.
  82. 82. Nelson KA, Moin NS, Bernhard AE (2009) Archaeal diversity and the prevalence of Crenarchaeota in salt marsh sediments. Appl Environ Microbiol 75: 4211–4215.
  83. 83. Kunin V, Raes J, Harris JK, Spear JR, Walker JJ, et al. (2008) Millimeter-scale genetic gradients and community-level molecular convergence in a hypersaline microbial mat. Mol Syst Biol 4: 198.
  84. 84. Shaw AK, Halpern AL, Beeson K, Tran B, Venter JC, et al. (2008) It’s all relative: ranking the diversity of aquatic bacterial communities. Environ Microbiol 10: 2200–2210.
  85. 85. Yousuf B, Sanadhya P, Keshri J, Jha B (2012) Comparative molecular analysis of chemolithoautotrophic bacterial diversity and community structure from coastal saline soils, Gujarat, India. BMC Microbiol 12: 150.