The microbial rare biosphere represents the largest pool of biodiversity on Earth and constitutes, in sum of all its members, a considerable part of a habitat’s biomass. Dormancy or starvation is typically used to explain the persistence of low-abundance microorganisms in the environment. We show that a low-abundance microorganism can be highly transcriptionally active while remaining in a zero-growth state for at least 7 weeks. Our results provide evidence that this zero growth at a high cellular activity state is driven by maintenance requirements. We show that this is true for a microbial keystone species, in particular a cosmopolitan but permanently low-abundance sulfate-reducing microorganism in wetlands that is involved in counterbalancing greenhouse gas emissions. In summary, our results provide an important step forward in understanding time-resolved activities of rare biosphere members relevant for ecosystem functions.
KEYWORDS: cryptic sulfur cycle, growth arrest, keystone species, maintenance, metatranscriptome, peatland
ABSTRACT
Microbial diversity in the environment is mainly concealed within the rare biosphere (all species with <0.1% relative abundance). While dormancy explains a low-abundance state very well, the mechanisms leading to rare but active microorganisms remain elusive. We used environmental systems biology to genomically and transcriptionally characterize “Candidatus Desulfosporosinus infrequens,” a low-abundance sulfate-reducing microorganism cosmopolitan to freshwater wetlands, where it contributes to cryptic sulfur cycling. We obtained its near-complete genome by metagenomics of acidic peat soil. In addition, we analyzed anoxic peat soil incubated under in situ-like conditions for 50 days by Desulfosporosinus-targeted qPCR and metatranscriptomics. The Desulfosporosinus population stayed at a constant low abundance under all incubation conditions, averaging 1.2 × 106 16S rRNA gene copies per cm³ soil. In contrast, transcriptional activity of “Ca. Desulfosporosinus infrequens” increased at day 36 by 56- to 188-fold when minor amendments of acetate, propionate, lactate, or butyrate were provided with sulfate, compared to the no-substrate-control. Overall transcriptional activity was driven by expression of genes encoding ribosomal proteins, energy metabolism, and stress response but not by expression of genes encoding cell growth-associated processes. Since our results did not support growth of these highly active microorganisms in terms of biomass increase or cell division, they had to invest their sole energy for maintenance, most likely counterbalancing acidic pH conditions. This finding explains how a rare biosphere member can contribute to a biogeochemically relevant process while remaining in a zero-growth state over a period of 50 days.
INTRODUCTION
The vast majority of microbial diversity worldwide is represented by the rare biosphere (1–4). This entity of microorganisms consists of all microbial species that have an arbitrarily defined relative population size of <0.1% in a given habitat at a given time (1–4). The rare biosphere is opposed by a much smaller number of moderately abundant or very abundant microbial species (≥0.1% and ≥1.0% relative abundance, respectively) (5), which are thought to be responsible for the major carbon and energy flow through a habitat as based on their cumulative biomass. However, there is accumulating experimental evidence that the rare biosphere is not just a “seed bank” of microorganisms that are waiting to become active and numerically dominant upon environmental change (3, 6) but also harbors metabolically active microorganisms with important ecosystem functions (4).
First hints for metabolically active rare biosphere members were evident from seasonal patterns of marine bacterioplankton species. Here, many taxa that displayed recurring annual abundance changes were of low abundance and even during their bloom periods never reached numerically abundant population sizes (7–9). In soil environments, removal of low-abundance species by dilution-to-extinction had a positive effect on intruding species, suggesting that active low-abundance species preoccupy ecological niches and thus slow down invasion (10–12). Soil microorganisms of low relative abundance were also shown to play a role in community-wide species interactions, e.g., by being involved in the production of antifungal compounds that protect plants from pathogens (13) or by constituting the core of microorganisms that respond to the presence of a particular plant species (14). Other examples include microorganisms with a specialized metabolism that sustain stable low-abundance populations in an ecosystem (3). For example, N2-fixing microorganisms in the ocean (15) or sulfate-reducing microorganisms (SRM) in peatlands (5, 16, 17) were shown to fulfill such key functions.
A peatland Desulfosporosinus species was one of the first examples identified as an active rare biosphere member contributing to an important ecosystem function (16). This SRM is involved in the cryptic sulfur cycle of peatlands (5, 16), which in turn controls the emission of the greenhouse gas CH4 from these globally relevant environments (17). Although porewater sulfate concentrations are typically quite low in peatlands (<300 µM) (17), these environments are characterized by temporally fluctuating high sulfate reduction rates (up to 1,800 nmol cm−3 day−1) (17). These rates can be in the same range as in sulfate-rich marine surface sediments, where sulfate reduction is one of the major anaerobic carbon degradation pathways (18, 19). In low-sulfate peatlands, such high sulfate reduction rates can be maintained only by rapid aerobic or anaerobic reoxidation of reduced sulfur species back to sulfate (17). Since SRM generally outcompete methanogens and syntrophically associated fermenters (20), they exert an important intrinsic control function on peatland CH4 production (21–23). This is important, since natural wetlands, such as peatlands, are estimated to be responsible for 30% of the annual emission of this potent greenhouse gas (24–26).
Little is known about the ecophysiology of metabolically active but low-abundance microorganisms. This lack of knowledge is clearly founded in their low numerical abundance, making it inherently difficult to study their metabolic responses or even to retrieve their genomes directly from the environment. In a preceding study, we could show that the low-abundance peatland Desulfosporosinus species mentioned above follows an ecological strategy to increase its cellular ribosome content while maintaining a stable low-abundance population size when exposed to favorable, sulfate-reducing conditions (5). This was unexpected since metabolic activity in bacteria and archaea is typically followed by growth (in terms of cell division or biomass increase) if they are not severely energy or nutrient limited (27) or engaged in major maintenance processes coping with (environmental) stress (28). The studied Desulfosporosinus species is found worldwide in a wide range of low-sulfate wetlands, including peatlands, permafrost soils, and rice paddy fields (5). This emphasizes its importance as a model organism for active rare biosphere members. In this study, we used an environmental systems biology approach to deepen our understanding of the cellular biology of this rare biosphere member. In particular, we retrieved its genome by metagenomics from native and incubated peat soil and followed its transcriptional responses in peat soil microcosms simulating transient in situ conditions, i.e., sulfate-reducing or methanogenic conditions in the background of minor amendments of the organic carbon degradation intermediate formate, acetate, propionate, lactate, or butyrate.
RESULTS
A near-complete genome of a rare biosphere member from peat soil.
We obtained the population genome of the low-abundance Desulfosporosinus species by coassembly and differential coverage binning of metagenomes obtained from native peat soil and 13C-labeled fractions of a DNA-stable isotope probing experiment of the same peatland (see Fig. S1 in the supplemental material) (29). The high-quality metagenome-assembled genome (MAG) SbF1 had a size of 5.3 Mbp (on 971 scaffolds), a G+C content of 42.6%, a CheckM-estimated completeness of 98.0%, a potential residual contamination of 3.9%, and 10% strain heterogeneity. Besides 16S and 23S rRNA genes, SbF1 carried 6,440 protein-coding genes (CDS), five 5S rRNA gene copies, 59 tRNAs, and 37 other ncRNAs, making a total of 6,543 predicted genomic features. The genome size and G+C content were in the same range as observed for genomes of cultured Desulfosporosinus species (3.0 to 5.9 Mbp and 42 to 44%, respectively) (30–34). Scaffolds containing rRNA genes had a higher coverage than the average coverage of all scaffolds (Fig. S1), indicating multiple rrn operon copies, as is known from genomes of other Desulfosporosinus species (on average, 9.3 rrn operons; range, 8 to 11) (35).
16S rRNA-based phylogenetic tree reconstruction placed SbF1 into a well-supported clade together with Desulfosporosinus sp. strain 44a-T3a (98.3% sequence identity), Desulfosporosinus sp. strain OT (98.8%), and Desulfosporosinus sp. strain 5apy (98.1%). The most similar validly described species was Desulfosporosinus lacus with a sequence identity of 97.5% (Fig. S2a). Phylogenomics confirmed Desulfosporosinus sp. OT as the closest relative (Fig. S2b) with average amino and nucleic acid identities (AAI and ANI, respectively) of 77% and 79%, respectively (Fig. S3). The intragenus AAI variability of Desulfosporosinus species was 69 to 93% (Fig. S3). Therefore, MAG SbF1 represents a novel species in this genus based on species-level thresholds of 99% for the 16S rRNA gene (36) and 96.5% for ANI (37).
The versatile energy metabolism of the low-abundance Desulfosporosinus.
Desulfosporosinus sp. MAG SbF1 encoded the complete canonical pathway for dissimilatory sulfate reduction (Fig. 1 and Table S1a). This encompassed sulfate adenylyltransferase (Sat), adenylyl-sulfate reductase (AprBA), dissimilatory sulfite reductase (DsrAB), and sulfide-releasing DsrC, which are sequentially involved in the reduction of sulfate to sulfide. In addition, genes encoding the electron-transferring QmoAB and DsrMKJOP complexes were detected, with their subunit composition being typical for Desulfosporosinus species (30, 31, 33, 34). Other dsr genes included dsrD, dsrN, and dsrT (38), with hitherto-unvalidated function; fdxD, which encodes a [4Fe4S]-ferredoxin; and a second set of DsrMK family-encoding genes (dsrM2 and dsrK2). SbF1 also encoded the trimeric dissimilatory sulfite reductase AsrABC (anaerobic sulfite reductase) (39).
SbF1 carried genes to utilize lactate, propionate, acetate, formate, or H2 as electron donors (Fig. 1). All enzymes necessary for propionate oxidation to the central metabolite pyruvate (including those belonging to a partial citric acid cycle) were encoded on two scaffolds (Table S1a). For lactate utilization, SbF1 carried three paralogs of glycolate/d-lactate/l-lactate dehydrogenase family genes (lutDF/glcDF). However, the substrate specificity of the encoded enzymes could not be inferred from sequence information alone. The transcription of lutDF and lutD_2 was stimulated by the addition of l-lactate (Fig. 1), which indicates that these genes encode functional lactate dehydrogenases (LDHs). The third paralog (glcDF, Table S1a) was not stimulated by lactate. LutDF was organized in an operon with a lactate permease (LutP) and a lactate regulatory gene (lutR). LutD_2 was organized in an operon with an electron-transferring flavoprotein (EtfBA_2), which resembled the electron-confurcating LDH/Etf complex in Acetobacterium woodii (40). LDHs have been shown to utilize both l- and d-lactate (40, 41). However, SbF1 also encoded a lactate racemase (LarA) and a lactate racemase-activating system (LarEBC) for interconversion of both stereoisomers (42).
Pyruvate, the intermediate product in propionate and lactate degradation, can be further oxidized to acetyl-CoA with either one of several pyruvate-ferredoxin oxidoreductases (PfoA) or formate C-acetyltransferase (PflD). Acetyl-CoA can then be completely oxidized to CO2 via the Wood-Ljungdahl pathway (43), which is complete in SbF1 (Fig. 1; Table S1a) and present in the genomes of all other sequenced Desulfosporosinus species (30, 31, 33, 34). Alternatively, acetyl-CoA may be incompletely oxidized to acetate via acetyl-phosphate by phosphate acetyltransferase (Pta) and acetate kinase (AckA). Pta and AckA are bidirectional enzymes, opening the possibility that acetate could be degraded via these two enzymes and the downstream Wood-Ljungdahl pathway to CO2.
Formate and H2 represented additional potential electron donors for SbF1. Its genome encoded three formate dehydrogenases (FDH). FDH-1 consists of three subunits (fdhCBA), while FDH-2 (FdhA_2) and FDH-3 (FdhA_3) are monomeric enzymes. In addition, [NiFe] hydrogenases of group 1 and 4f, as well as [FeFe] hydrogenases of group A (44), were encoded. Homologs of genes for butyrate oxidation were missing in SbF1 (45), which is in contrast to other Desulfosporosinus species (e.g., Desulfosporosinus orientis). Both glycolysis and gluconeogenesis were complete. However, neither a glucokinase nor a phosphotransferase system (PTS) was found. Coupling of electron transfer to energy conservation could be mediated in SbF1 by a H+/Na+-pumping Rnf complex (RnfCDGEAB) (46) and an NADH dehydrogenase (respiratory complex I, NuoABCDEFGHIJKLMN). In addition, the complete gene set for ATP synthase (AtpABCDEFGH) was identified (Fig. 1; Table S1a).
Long-term transcriptional activity of Desulfosporosinus sp. MAG SbF1 at zero growth.
Naturally occurring hot spots of sulfate-reducing activity in peat soil (47–50) were mimicked by periodically amending sulfate in the lower- micromolar range to anoxic peat microcosms (every 3 to 7 days) and comparing this to unamended (i.e., methanogenic) control microcosms. In addition, sulfate-reducing and methanogenic microcosms received, in triplicates, periodic amendments of either formate, acetate, propionate, lactate, or butyrate compared to controls without amendment. Substrate supply did generally not exceed 100 to 200 µM, thus again mimicking in situ concentrations of these naturally occurring organic carbon degradation intermediates in peatlands (5). The overall Desulfosporosinus population remained stable throughout the incubation period in the various microcosms (on average, 1.2 × 106 16S rRNA gene copies per cm3 of soil [Fig. 2a]). Compared to the total bacterial and archaeal community, this resembled a relative abundance of 0.018% when corrected for the average 9.3 rrn operons per genome in the genus Desulfosporosinus (35). The 16S rRNA gene of Desulfosporosinus sp. MAG SbF1 was 100% identical to OTU0051, which dominated the Desulfosporosinus population as evident from a previously published 16S rRNA (gene) amplicon survey of the same microcosms (74% of all Desulfosporosinus amplicons) (5). In contrast to its stable low abundance, the overall Desulfosporosinus population substantially increased its 16S rRNA copy numbers by 2.2, 4.9, 5.9, or 13.6-fold in sulfate-reducing incubations stimulated by either acetate, propionate, lactate, or butyrate, respectively. In contrast, Desulfosporosinus 16S rRNA copy numbers remained stable or even slightly decreased in the sulfate-amended no-substrate control and the methanogenic incubations (Fig. S4) (5). Again, these increases were mainly reflected in changes of OTU0051 (Desulfosporosinus sp. MAG SbF1) as shown in the amplicon study mentioned above (5).
We used metatranscriptomics of the same microcosms to analyze whether this strong increase in 16S rRNA copies at zero growth was accompanied by gene expression of metabolic pathways and cell growth-associated processes in Desulfosporosinus sp. MAG SbF1. Compared to the initial soil, the overall transcriptional activity of SbF1 steadily increased at days 8 and 36 in sulfate-reducing incubations stimulated by either acetate, propionate, lactate, or butyrate. In contrast, all methanogenic incubations as well as the sulfate-reducing formate and no-substrate incubations showed, after an initial stimulation until day 8, a steady or even mildly decreasing overall transcriptional activity (Fig. 2b). At day 36, normalized mRNA counts of SbF1 were 56-, 80-, 62-, or 188-fold higher in sulfate-reducing incubations stimulated by either acetate, propionate, lactate, or butyrate, respectively, compared to the no-substrate control and constituted between 0.11 ± 0.13% (acetate) and 0.36 ± 0.02% (butyrate) of all transcripts in the corresponding metatranscriptomes (Fig. 2b). This substrate-specific activity was driven by the increased transcription of genes encoding ribosomal proteins as general activity markers (Fig. 3; Table S1a) and energy metabolism genes, including all canonical dissimilatory sulfate reduction genes (Fig. 4; Table S1a). For example, Spearman’s rank correlation coefficients of normalized dsrA and dsrB transcript counts compared to the sum of normalized SbF1 mRNA counts were 0.91 and 0.90, respectively (FDR-adjusted P value < 0.001). Normalized transcript counts of other enzyme complexes involved in the central metabolism of SbF1, such as the ATP synthase, the NADH dehydrogenase (complex I), and ribosomal proteins, followed the same transcriptional pattern (Fig. 4; Table S1a) with an average Spearman rank correlation coefficient of 0.79 ± 0.07 (n = 72, FDR-adjusted P value < 0.05) to the sum of normalized SbF1 mRNA counts. Interestingly, transcription of genes encoding proteins involved in general stress response was stimulated as well. In particular, genes encoding the universal stress promoter UspA, the GroES/GroEL and DnaK chaperons, and the proteolytic subunit of ATP-dependent Clp protease (ClpP) showed an increased transcription (Fig. 4) with an average Spearman rank correlation coefficient of 0.76 ± 0.04 (n = 5, FDR-adjusted P value < 0.05) to the sum of normalized SbF1 mRNA counts. In the microcosms with stimulated transcriptional activity, dsrA transcripts of SbF1 constituted also a dominant fraction of overall dsrA transcripts ranging from 11% to 53% at day 36 (Fig. S6).
To evaluate whether a hidden turnover of biomass (cryptic growth) was underlying the stable Desulfosporosinus population, we screened COG categories D, L, and M for expression of indicator genes that encode functions in cell division (e.g., ftsZ or minE), DNA replication (e.g., gyrBA, dnaC, and dnaG), and cell envelope biogenesis (e.g., murABCDEFGI), respectively. Genes that unambiguously encoded such functions (Table S1a) showed only very minor or no increases in transcripts over time (Fig. 3; detailed in Fig. S5). Extension of this analysis to all genes belonging to COG D (n = 73), L (n = 280), and M (n = 215) showed that the average Spearman rank correlation coefficient to the sum of normalized mRNA counts was only 0.45 ± 0.13 (FDR-adjusted P value < 0.05 [Table S1a]).
We also analyzed genes reported to be upregulated immediately after phage infection, which is an important ecological control of bacterial population size. Respective genes in Bacillus subtilis encode, e.g., functions in DNA and protein metabolism and include the ribonucleoside-diphosphate reductase (nrdEF) and aspartyl/glutamyl-tRNA amidotransferase (gatCAB) (51). However, homologs in SbF1 did not show increased expression in the incubations with increased total transcriptional activity (Table S1a). This was reflected in an average Spearman rank correlation coefficient of only 0.60 ± 0.06 (n = 4, FDR-adjusted P value < 0.05) to the sum of normalized SbF1 mRNA counts. The same was true when screening for active sporulation of a Desulfosporosinus subpopulation as an alternative explanation for a stable low-abundance population. The identified sporulation genes (spo0A to spoVT) did not show any prominent increase in transcript numbers as well, with the only exception of spoIIIAD, which was stimulated in propionate- and sulfate-amended microcosms (Table S1a). Again, expression of genes involved in sporulation had a low average Spearman rank correlation coefficient of 0.44 ± 0.13 (n = 22, FDR-adjusted P value < 0.05) to the sum of normalized SbF1 mRNA counts.
The individual incubation regimes triggered in addition transcriptional activation of the respective substrate degradation pathways of Desulfosporosinus sp. MAG SbF1. For example, all genes necessary for the conversion of propionate to pyruvate were overexpressed only upon addition of propionate and sulfate but not in any other incubation type. The same was true for lactate degradation, where genes encoding the lactate permease, lactate racemase, and two of the detected lactate dehydrogenases were overexpressed upon addition of both lactate and sulfate but not in incubations with lactate only (Fig. 4). Although genes encoding phosphotransacetylase and acetate kinase were overexpressed under lactate and propionate, the complete Wood-Ljungdahl pathway was overexpressed as well, which indicates that at least part of these substrates were completely degraded to CO2 rather than to acetate and CO2. This conclusion was supported by the overexpression of the Wood-Ljungdahl pathway in incubations amended with acetate and sulfate. Interestingly, the Wood-Ljungdahl pathway was also overexpressed upon addition of butyrate and sulfate. Under such conditions, Desulfosporosinus sp. MAG SbF1 apparently relies on acetate released by a butyrate utilizer, as it lacks the capability for butyrate oxidation, albeit failed recovery of the butyrate degradation pathway during binning cannot be excluded.
DISCUSSION
Current knowledge on the interconnection of energy metabolism, gene expression, cell division, and population growth in microorganisms is mainly based on pure cultures that are maintained in the laboratory. Under ideal conditions, a single Escherichia coli cell would grow to a population with the mass of the Earth within 2 days. Clearly, this does not occur, but the discrepancy between potential and actual growth underscores that microorganisms spend the vast majority of their time not dividing (27, 52). A large fraction of these microorganisms are part of the rare biosphere. For example, in the studied peatland, the sum of all low-abundance species made up approximately 12% of the total bacterial and archaeal 16S rRNA genes (5). In other soils, low-abundance Alphaproteobacteria and Bacteroidetes alone constituted in sum 10% and 9% of the total bacterial population, respectively, while all low-abundance populations summed up to 37% of all bacteria (14). Upon strong environmental change, low-abundance microorganisms often grow to numerically abundant populations and replace dominant species as observed for microbial community changes after an oil spill (53, 54) or in the response of soil microorganisms toward the presence of plants (14). However, subtle environmental changes (5) or recurring seasonal shifts (7, 9, 55) often lead to rather small shifts in low-abundance populations without rare biosphere members becoming numerically dominant.
The low-abundance Desulfosporosinus sp. MAG SbF1 represents an interesting case of the latter response type. When exposed to favorable, sulfate-reducing conditions in peat soil microcosms, the overall Desulfosporosinus population did not increase its population size of about 1.2 × 106 16S rRNA gene copies cm−3 soil (Fig. 2a) but strongly increased its cellular ribosome content by up to 1 order of magnitude (see Fig. S4 in the supplemental material) (5). In a preceding 16S rRNA (gene) amplicon study which analyzed the same microcosms, we could show that Desulfosporosinus OTU0051 is the major constituent of this Desulfosporosinus population (74% of all Desulfosporosinus amplicons) and correlated best in its 16S rRNA response to sulfate turnover among all identified SRM (5). Here, we reanalyzed these microcosms to expand upon this observation by genome-centric metatranscriptomics and to test whether the increase in cellular ribosome content is indeed translated into transcriptional and, as a consequence, metabolic activity. Desulfosporosinus OTU0051 was 100% identical to the 16S rRNA gene of Desulfosporosinus sp. MAG SbF1, which was retrieved in this study and as such represented the major Desulfosporosinus population. In support of this conclusion, increases in 16S rRNA copies of the overall Desulfosporosinus population (Fig. S4) (5) clearly corresponded to increased transcription of genes coding for ribosomal proteins in Desulfosporosinus sp. MAG SbF1 (Fig. 3; Table S1a) (5). This cellular ribosome increase under sulfate-reducing conditions correlated well with an increase in all normalized mRNA counts (Fig. 2b). This is the first time that changes in population-wide 16S rRNA levels are proven to be directly linked to transcriptional activity for a rare biosphere member.
Analyzing the transcriptional response of a rare biosphere member under in situ-like conditions opens the unique opportunity to gain insights into its ecophysiology. Desulfosporosinus sp. MAG SbF1 clearly overexpressed its sulfate reduction pathway under sulfate amendment when supplied with either acetate, lactate, propionate, or butyrate compared to the no-substrate and the methanogenic controls (Fig. 4). Here, dsrA transcripts of Desulfosporosinus sp. MAG SbF1 represented a dominant fraction of all detected dsrA transcripts (up to 53%), supporting its importance for the observed sulfate turnover in the analyzed microcosms (Fig. S6). Detailed analysis of the transcribed carbon degradation pathways showed that Desulfosporosinus sp. MAG SbF1 is able to oxidize propionate, lactate, and acetate completely to CO2. This was unexpected, since all described species of the genus Desulfosporosinus are so far known as incomplete oxidizers of organic compounds to acetate and CO2 (56–59). Under butyrate-amended conditions, Desulfosporosinus sp. MAG SbF1 presumably relied on uptake of acetate supplied by a primary butyrate oxidizer. This shows that Desulfosporosinus sp. MAG SbF1 is capable of utilizing diverse substrates that represent the most important carbon degradation intermediates measured in peatlands (60, 61).
The question remains which mechanisms are at work that keep the transcriptionally active Desulfosporosinus sp. MAG SbF1 population in a stable low-abundance state. Ongoing growth could be hidden by continuous predation, viral lysis, or active sporulation of a major subpopulation. To answer this question, we analyzed expression patterns of genes involved in cell growth-associated processes. Compared to the strong overexpression of metabolic or ribosomal protein genes, transcription of genes essential for DNA replication, cell division, and cell envelope biogenesis did not increase or increased only marginally (Fig. 3; Fig. S5). In contrast, retentostat studies on cultured Firmicutes held in a (near-) zero-growth state revealed that expression of genes involved in cell growth, central energy metabolism, and the translational apparatus was always coregulated, showing either a joint increased expression in Bacillus subtilis (62) or an invariable expression in Lactobacillus plantarum (63) when comparing active growth to (near-) zero growth. In addition, there is experimental evidence that in the lag phase of batch cultures, i.e., in the transition from no growth to growth, transcription of growth-related genes is not stable but increases due to the overall activation of cellular processes (64). In this context, the lack of an increasing transcription of growth-related genes would clearly indicate a state of (near-) zero growth rather than an actively dividing population that is kept stable by an equally high growth and mortality or sporulation rate. This conclusion is further supported by the lack of overexpressed sporulation genes or genes upregulated directly after phage attack (Table S1a and c).
Nevertheless, the ATP generated by the induced energy metabolism has to be consumed. If not used for growth, it has to be invested completely for maintenance according to the Herbert-Pirt relation qs = ms + µ/Ysxmax, where qs is the biomass-specific consumption rate, ms is the maintenance coefficient, µ is the specific growth rate, and Ysxmax is the maximum growth yield (65, 66). Based on the concept of a species-independent maintenance energy requirement as laid out by reference 67, and further developed by reference 28, it can be calculated that Desulfosporosinus sp. MAG SbF1 would need to consume 1.5 fmol sulfate per day to maintain a single cell in our incubations when, e.g., incompletely oxidizing lactate to acetate (detailed in Text S1 in the supplemental material). This is in agreement with experimentally determined maintenance requirements of Desulfotomaculum putei (68) but 2 orders of magnitude smaller than the cell-specific sulfate reduction rates of Desulfosporosinus sp. MAG SbF1 estimated previously in a similar experimental setup of the same peat soil by reference 16 (here the responsive but low-abundance Desulfosporosinus OTU was 99.8% identical to the 16S rRNA gene of Desulfosporosinus sp. MAG SbF1). However, maintenance requirements are known to increase upon production of storage compounds or to counterbalance environmental stress (28). We found no indication for the former scenario but observed overexpression of the universal stress promoter UspA, which is one of the most abundant proteins in growth-arrested cells (69). In addition, we observed overexpression of the chaperons GroES/GroEL and DnaK and of the protease ClpP, which were all previously linked to low-pH stress response at the expense of ATP consumption (70–74). Since the pH in the analyzed peat soil incubations varied between 4.1 and 5.0 (5), coping with a low pH would be the most parsimonious explanation for increased maintenance requirements. In this context, one may speculate whether the overexpressed ATP synthase might have operated as an ATPase to pump protons out of the cell at the expense of ATP hydrolysis, which is a known response mechanisms toward mildly acidic pH (74). Similarly, the overexpressed sulfate reduction pathway, including complex I and the membrane quinone shuttle, might have been co-utilized as a proton pump without harvesting the membrane potential for ATP generation. Since active sulfate reduction would also consume protons in the vicinity of Desulfosporosinus sp. MAG SbF1 and thus slowly increase its surrounding pH, a high metabolic activity at concomitant zero growth controlled by maintenance requirements would make sense. Favorable conditions lasting longer than the 50 days studied here may result in increasing the surrounding pH enough for surplus energy to be invested in growth. This would explain the observed abundance increases of Desulfosporosinus 16S rRNA genes after 73 days of incubation in a previous study (16) and the detectable 13C incorporation into the genome of the rare peatland Desulfosporosinus after 120 days of incubations (increased coverage in the DNA-SIP metagenome). In the context of the dynamic occurrence of sulfate reduction in peatlands, e.g., driven by changes in the water table as the oxic-anoxic interface or the complex flow paths of infiltrating and exfiltrating water that create distinct spatial and temporal activity patterns (47–50), the studied Desulfosporosinus population may experience time periods of favorable conditions (as in our incubations) that would result first in activity at zero growth to counteract (pH) stress and, if provided long enough, eventually also in growth.
Our results are important in the context of the increasing awareness that the microbial rare biosphere is not only the largest pool of biodiversity on Earth (1–4) but in sum of all its low-abundance members constitutes also a large part of the biomass in a given habitat (5, 14). Understanding the mechanisms governing this low-abundance prevalence and its direct impact on ecosystem functions and biogeochemical cycling is thus of utmost importance. Desulfosporosinus sp. MAG SbF1 has been repeatedly shown to be involved in cryptic sulfur cycling in peatlands (5, 16)—a process that counterbalances the emission of the greenhouse gas methane due to the competitive advantage of SRM compared to microorganisms involved in the methanogenic degradation pathways (20). This species can be found worldwide in low-sulfate environments impacted by cryptic sulfur cycling, including not only peatlands but also permafrost soils, rice paddies, and other wetland types (5). Here, we provided proof that Desulfosporosinus sp. MAG SbF1 is indeed involved in the degradation of important anaerobic carbon degradation intermediates in peatlands while sustaining a low-abundance population. It has a generalist lifestyle in respect to the usable carbon sources, re-emphasizing its importance in the carbon and sulfur cycle of peatlands. Our results provide an important step forward in understanding the microbial ecology of biogeochemically relevant microorganisms and show that low-abundance keystone species can be studied “in the wild” using modern environmental systems biology approaches.
Proposal of “Candidatus Desulfosporosinus infrequens.”
Based on its phylogenetic placement and novel ecophysiological behavior, we propose that Desulfosporosinus sp. MAG SbF1 represents a novel species with the provisional name “Candidatus Desulfosporosinus infrequens” sp. nov. (in.fre′quens. L. adj. infrequens, rare, referring to its low relative abundance). Based on its genome-derived metabolic potential and support from metatranscriptomics, “Ca. Desulfosporosinus infrequens” is capable of complete oxidation of acetate, propionate, and lactate with sulfate as the electron acceptor, with further potential for oxidation of molecular hydrogen (Fig. 1).
MATERIALS AND METHODS
Genome assembly, binning, and phylogenetic inference.
Sampling of peat soil from the acidic peatland Schlöppnerbrunnen II (Germany), DNA-stable isotope probing (DNA-SIP), total nucleic acid extraction, metagenome sequencing and assembly, and coverage-based binning were described previously (5, 16, 29). In brief, DNA from native peat soil (10- to 20-cm depth) and DNA pooled from 16 13C-enriched fractions (density, 1.715 to 1.726 g ml−1) of a previous DNA-SIP experiment with soil from the same site (16) was sequenced using the Illumina HiSeq 2000 system. DNA-SIP was performed after a 120-day incubation (again, 10- to 20-cm depth) that was periodically amended with small dosages of sulfate and first a mixture of unlabeled formate, acetate, propionate, and lactate for 2 weeks and thereafter a mixture of 13C-labeled formate, acetate, propionate, and lactate (all in the lower-micromolar range) (16). Raw reads were quality filtered, trimmed, and coassembled (native soil, 385 million reads; DNA-SIP, 576 million reads) using the CLC Genomics Workbench 5.5.1 (CLC Bio). Differential coverage binning was applied to extract the Desulfosporosinus metagenome-assembled genome (MAG) (75). As expected (16), the Desulfosporosinus MAG was of low abundance in the native soil with an average coverage of 0.026 while enriched in the SIP sample with an average coverage of 34 (detailed per scaffold in Table S1b). A side effect of sequencing a DNA-SIP sample is an apparent G+C content skew, which was normalized arbitrarily to improve binning using the following formula (29, 76): (coverage/G+C content9) × 1015.
Scaffolds containing the 16S and 23S rRNA genes were successfully identified using paired-end linkage data (75). Completeness, contamination, and strain heterogeneity were estimated using CheckM 1.0.6 (77).
Phylogenomic analysis of the Desulfosporosinus MAG was based on a concatenated set of 34 phylogenetically informative marker genes as defined by reference 77 and the Bayesian phylogeny inference method PhyloBayes using the CAT-GTR model (78). 16S rRNA gene-based phylogeny was inferred using the ARB SILVA database r126 as a reference (79), the SINA aligner (80), and the substitution model testing and maximum likelihood treeing method IQ-TREE (81). Pairwise 16S rRNA gene sequence identities were calculated with T-Coffee 11 (82). Pairwise average nucleic and amino acid identities (ANI and AAI, respectively) (37) between protein-encoding genes of the Desulfosporosinus MAG and reference genomes were calculated as described previously (29).
Genome annotation.
The genome was annotated using the MicroScope annotation platform (83). Annotation refinement for selected genes was done as follows: proteins with an amino acid identity ≥40% (over ≥80% of the sequence) to a Swiss-Prot entry (84) or a curated MaGe annotation (83) or proteins described in the literature were annotated as true homologs of known proteins. The same was true if classification according to InterPro families (85, 86), TIGRFAMs (87), and/or FIGfams (88) led to an unambiguous annotation. Proteins with an amino acid identity ≥25% (over ≥80% of the sequence) to a Swiss-Prot or TrEMBL (84) entry were annotated as putative homologs of the respective database entries. In addition, classification according to COG (89) or InterPro superfamilies, domains, or binding sites was used to call putative homologs in cases of an unambiguous annotation. Membership to syntenic regions (operons) was considered additional support to call true or putative homologs.
Metatranscriptomics from single-substrate incubations.
Metatranscriptomic data sets of anoxic peat soil slurry microcosms that were previously described (29) were re-analyzed in the context of published Desulfosporosinus qPCR data (5) of the same microcosms. In brief, anoxic microcosms were incubated at 14°C in the dark for 50 days and regularly either amended with small amounts of sulfate (76 to 387 µM final concentrations) or incubated without an external electron acceptor. Formate, acetate, propionate, lactate, butyrate (<200 µM), or no external electron donor was added to biological triplicates each. DNA and/or RNA wase extracted from the native soil and after 5, 8, 15, 26, 36, and 50 days of incubations. Quantitative PCR data describing 16S rRNA gene copies of the complete Desulfosporosinus population (three species-level operational taxonomic units [OTUs]: 74% OTU0051, 26% OTU0228, and <0.001% OTU7382; average of the native peat soil and all microcosms) in comparison to the overall bacterial and archaeal community were taken from reference 5 and used in this study to put the metatranscriptome data into the perspective of population dynamics. PCR conditions are given in reference 5. Metatranscriptome sequencing was done from each of the biological replicates using the Illumina HiSeq 2000/2500 platform (27 to 188 million reads per sample). Raw reads were quality filtered as described previously (29) and mapped to the Desulfosporosinus MAG in a background of all other metagenome-assembled scaffolds using Bowtie 2 at default settings (90). Counting of mapped reads to protein-encoding genes (CDS) was performed with featureCounts 1.5.0 (91).
Statistical analysis of Desulfosporosinus-specific transcripts.
Counts of mapped transcript reads were normalized to the length of the respective gene and the sequencing depth of the respective metatranscriptome, resulting in FPKM (fragments per kilobase per million total fragments) values. Thereafter, we used an unsupervised approach to identify CDS expression stimulated by sulfate and the different substrates regimes. First, we applied the DESeq2 R package (92, 93) to identify differentially expressed CDS. Treatments without external sulfate added and samples after 8 days of incubations had too few transcript counts to be used for a statistical approach. Therefore, we limited our analysis to pairwise comparison of sulfate-stimulated microcosms after 36 days of incubation. We compared each substrate regime to the no-substrate controls and each other. The set of all significantly differentially expressed CDS (FDR-adjusted P value < 0.05) was further clustered into response groups. For clustering, we calculated pairwise Pearson correlation coefficients (r) of variance stabilized counts (cor function in R) and transformed this into distances (1 – r), followed by hierarchical clustering (hclust function in R). Variance stabilization was performed using the rlog function of the DESeq2 package. Spearman’s rank correlation of FPKM values for each gene to the total relative mRNA counts was performed with cor.test in R using the data from all treatments and replicates.
Sequence data availability.
The MAG SbF1 is available at MicroScope (https://www.genoscope.cns.fr/agc/microscope/) and is also deposited under the GenBank accession number OMOF01000000. Metagenome and transcriptomic data are available at the Joint Genome Institute (https://genome.jgi.doe.gov/) and are also deposited under the GenBank accession numbers PRJNA412436 and PRJNA412438, respectively.
ACKNOWLEDGMENTS
We are grateful to Mads Albertsen, Norbert Bittner, Tijana Glavina del Rio, Florian Goldenberg, Craig Herbold, Stephan Köstlbacher, Per H. Nielsen, Ulrich Stingl, and Susannah G. Tringe for sequence analysis and technical support. We further thank Bernhard Schink for help in naming “Ca. Desulfosporosinus infrequens”; Kenneth Wasmund for valuable feedback; and Johannes Wittmann, Jan-Ulrich Kreft, and Silvia Bulgheresi for helpful expert opinions.
We acknowledge the LABGeM (CEA/IG/Genoscope and& CNRS UMR8030) and the France Génomique National infrastructure (funded as part of Investissement d’avenir program managed by Agence Nationale pour la Recherche, contract ANR-10-INBS-09) for support with the MicroScope annotation platform. The work conducted by the Joint Genome Institute was supported by the Office of Science of the U.S. Department of Energy under contract no. DE-AC02-05CH11231. This research was financially supported by the Austrian Science Fund (FWF, P23117-B17 to M.P. and A.L. and P25111-B22 to A.L.), the U.S. Department of Energy (CSP605 to M.P. and A.L.), the German Research Foundation (DFG, PE 2147/1-1 to M.P.), and the European Union (FP7-People-2013-CIG, grant no. PCIG14-GA-2013-630188 to M.P.).
The authors declare no conflict of interest.
Footnotes
Citation Hausmann B, Pelikan C, Rattei T, Loy A, Pester M. 2019. Long-term transcriptional activity at zero growth of a cosmopolitan rare biosphere member. mBio 10:e02189-18. https://doi.org/10.1128/mBio.02189-18.
REFERENCES
- 1.Sogin ML, Morrison HG, Huber JA, Mark Welch D, Huse SM, Neal PR, Arrieta JM, Herndl GJ. 2006. Microbial diversity in the deep sea and the underexplored rare biosphere. Proc Natl Acad Sci U S A 103:12115–12120. doi: 10.1073/pnas.0605127103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Pedrós-Alió C. 2012. The rare bacterial biosphere. Annu Rev Mar Sci 4:449–466. doi: 10.1146/annurev-marine-120710-100948. [DOI] [PubMed] [Google Scholar]
- 3.Lynch MDJ, Neufeld JD. 2015. Ecology and exploration of the rare biosphere. Nat Rev Microbiol 13:217–229. doi: 10.1038/nrmicro3400. [DOI] [PubMed] [Google Scholar]
- 4.Jousset A, Bienhold C, Chatzinotas A, Gallien L, Gobet A, Kurm V, Küsel K, Rillig MC, Rivett DW, Salles JF, van der Heijden MGA, Youssef NH, Zhang X, Wei Z, Hol WHG. 2017. Where less may be more: how the rare biosphere pulls ecosystems strings. ISME J 11:853–862. doi: 10.1038/ismej.2016.174. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Hausmann B, Knorr K-H, Schreck K, Tringe SG, Glavina del Rio T, Loy A, Pester M. 2016. Consortia of low-abundance bacteria drive sulfate reduction-dependent degradation of fermentation products in peat soil microcosms. ISME J 10:2365–2375. doi: 10.1038/ismej.2016.42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Müller AL, de Rezende JR, Hubert CRJ, Kjeldsen KU, Lagkouvardos I, Berry D, Jørgensen BB, Loy A. 2014. Endospores of thermophilic bacteria as tracers of microbial dispersal by ocean currents. ISME J 8:1153–1165. doi: 10.1038/ismej.2013.225. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Campbell BJ, Yu L, Heidelberg JF, Kirchman DL. 2011. Activity of abundant and rare bacteria in a coastal ocean. Proc Natl Acad Sci U S A 108:12776–12781. doi: 10.1073/pnas.1101405108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Hugoni M, Taib N, Debroas D, Domaizon I, Jouan Dufournel I, Bronner G, Salter I, Agogué H, Mary I, Galand PE. 2013. Structure of the rare archaeal biosphere and seasonal dynamics of active ecotypes in surface coastal waters. Proc Natl Acad Sci U S A 110:6004–6009. doi: 10.1073/pnas.1216863110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Alonso-Sáez L, Díaz-Pérez L, Morán XAG. 2015. The hidden seasonality of the rare biosphere in coastal marine bacterioplankton. Environ Microbiol 17:3766–3780. doi: 10.1111/1462-2920.12801. [DOI] [PubMed] [Google Scholar]
- 10.van Elsas JD, Chiurazzi M, Mallon CA, Elhottova D, Kristufek V, Salles JF. 2012. Microbial diversity determines the invasion of soil by a bacterial pathogen. Proc Natl Acad Sci U S A 109:1159–1164. doi: 10.1073/pnas.1109326109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Vivant A-L, Garmyn D, Maron P-A, Nowak V, Piveteau P. 2013. Microbial diversity and structure are drivers of the biological barrier effect against Listeria monocytogenes in soil. PLoS One 8:e76991. doi: 10.1371/journal.pone.0076991. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Mallon CA, Poly F, Le Roux X, Marring I, van Elsas JD, Salles JF. 2015. Resource pulses can alleviate the biodiversity-invasion relationship in soil microbial communities. Ecology 96:915–926. doi: 10.1890/14-1001.1. [DOI] [PubMed] [Google Scholar]
- 13.Hol WHG, Garbeva P, Hordijk C, Hundscheid PJ, Gunnewiek PJAK, Van Agtmaal M, Kuramae EE, De Boer W. 2015. Non-random species loss in bacterial communities reduces antifungal volatile production. Ecology 96:2042–2048. doi: 10.1890/14-2359.1. [DOI] [PubMed] [Google Scholar]
- 14.Dawson W, Hör J, Egert M, van KM, Pester M. 2017. A small number of low-abundance bacteria dominate plant species-specific responses during rhizosphere colonization. Front Microbiol 8:975. doi: 10.3389/fmicb.2017.00975. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Großkopf T, Mohr W, Baustian T, Schunck H, Gill D, Kuypers MMM, Lavik G, Schmitz RA, Wallace DWR, LaRoche J. 2012. Doubling of marine dinitrogen-fixation rates based on direct measurements. Nature 488:361–364. doi: 10.1038/nature11338. [DOI] [PubMed] [Google Scholar]
- 16.Pester M, Bittner N, Deevong P, Wagner M, Loy A. 2010. A ‘rare biosphere’ microorganism contributes to sulfate reduction in a peatland. ISME J 4:1591–1602. doi: 10.1038/ismej.2010.75. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Pester M, Knorr K-H, Friedrich MW, Wagner M, Loy A. 2012. Sulfate-reducing microorganisms in wetlands—fameless actors in carbon cycling and climate change. Front Microbiol 3:72. doi: 10.3389/fmicb.2012.00072. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Jørgensen BB. 1982. Mineralization of organic matter in the sea bed—the role of sulphate reduction. Nature 296:643–645. doi: 10.1038/296643a0. [DOI] [Google Scholar]
- 19.Bowles MW, Mogollón JM, Kasten S, Zabel M, Hinrichs K-U. 2014. Global rates of marine sulfate reduction and implications for sub-sea-floor metabolic activities. Science 344:889–891. doi: 10.1126/science.1249213. [DOI] [PubMed] [Google Scholar]
- 20.Muyzer G, Stams AJM. 2008. The ecology and biotechnology of sulphate-reducing bacteria. Nat Rev Microbiol 6:441–454. doi: 10.1038/nrmicro1892. [DOI] [PubMed] [Google Scholar]
- 21.Gauci V, Matthews E, Dise N, Walter B, Koch D, Granberg G, Vile M. 2004. Sulfur pollution suppression of the wetland methane source in the 20th and 21st centuries. Proc Natl Acad Sci U S A 101:12583–12587. doi: 10.1073/pnas.0404412101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Gauci V, Dise N, Blake S. 2005. Long-term suppression of wetland methane flux following a pulse of simulated acid rain. Geophys Res Lett 32:L12804. [Google Scholar]
- 23.Gauci V, Chapman SJ. 2006. Simultaneous inhibition of CH4 efflux and stimulation of sulphate reduction in peat subject to simulated acid rain. Soil Biol Biochem 38:3506–3510. doi: 10.1016/j.soilbio.2006.05.011. [DOI] [Google Scholar]
- 24.Ciais P, Sabine C, Bala G, Bopp L, Brovkin V, Canadell J, Chhabra A, DeFries R, Galloway J, Heimann M, Jones C, Quéré CL, Myneni R, Piao S, Thornton P. 2013. Carbon and other biogeochemical cycles In Stocker T, Qin D, Plattner G-K, Tignor M, Allen S, Boschung J, Nauels A, Xia Y, Bex V, Midgley P (ed), Climate change 2013. The physical science basis. Cambridge University Press, Cambridge, United Kingdom. [Google Scholar]
- 25.Kirschke S, Bousquet P, Ciais P, Saunois M, Canadell JG, Dlugokencky EJ, Bergamaschi P, Bergmann D, Blake DR, Bruhwiler L, Cameron-Smith P, Castaldi S, Chevallier F, Feng L, Fraser A, Heimann M, Hodson EL, Houweling S, Josse B, Fraser PJ, Krummel PB, Lamarque J-F, Langenfelds RL, Le Quéré C, Naik V, O’Doherty S, Palmer PI, Pison I, Plummer D, Poulter B, Prinn RG, Rigby M, Ringeval B, Santini M, Schmidt M, Shindell DT, Simpson IJ, Spahni R, Steele LP, Strode SA, Sudo K, Szopa S, van der Werf GR, Voulgarakis A, van Weele M, Weiss RF, Williams JE, Zeng G. 2013. Three decades of global methane sources and sinks. Nat Geosci 6:813–823. doi: 10.1038/ngeo1955. [DOI] [Google Scholar]
- 26.Saunois M, Bousquet P, Poulter B, Peregon A, Ciais P, Canadell JG, Dlugokencky EJ, Etiope G, Bastviken D, Houweling S, Janssens-Maenhout G, Tubiello FN, Castaldi S, Jackson RB, Alexe M, Arora VK, Beerling DJ, Bergamaschi P, Blake DR, Brailsford G, Brovkin V, Bruhwiler L, Crevoisier C, Crill P, Covey K, Curry C, Frankenberg C, Gedney N, Höglund-Isaksson L, Ishizawa M, Ito A, Joos F, Kim H-S, Kleinen T, Krummel P, Lamarque J-F, Langenfelds R, Locatelli R, Machida T, Maksyutov S, McDonald KC, Marshall J, Melton JR, Morino I, Naik V, O’Doherty S, Parmentier F-JW, Patra PK, Peng C, Peng S, Peters GP, Pison I, Prigent C, Prinn R, Ramonet M, Riley WJ, Saito M, Santini M, Schroeder R, Simpson IJ, Spahni R, Steele P, Takizawa A, Thornton BF, Tian H, Tohjima Y, Viovy N, Voulgarakis A, van Weele M, van der Werf GR, Weiss R, Wiedinmyer C, Wilton DJ, Wiltshire A, Worthy D, Wunch D, Xu X, Yoshida Y, Zhang B, Zhang Z, Zhu Q. 2016. The global methane budget 2000–2012. Earth Syst Sci Data 8:697–751. [Google Scholar]
- 27.Lever MA, Rogers KL, Lloyd KG, Overmann J, Schink B, Thauer RK, Hoehler TM, Jørgensen BB. 2015. Life under extreme energy limitation: a synthesis of laboratory- and field-based investigations. FEMS Microbiol Rev 39:688–728. doi: 10.1093/femsre/fuv020. [DOI] [PubMed] [Google Scholar]
- 28.Harder J. 2006. Species-independent maintenance energy and natural population sizes. FEMS Microbiol Ecol 23:39–44. doi: 10.1111/j.1574-6941.1997.tb00389.x. [DOI] [Google Scholar]
- 29.Hausmann B, Pelikan C, Herbold CW, Köstlbacher S, Albertsen M, Eichorst SA, Glavina Del Rio T, Huemer M, Nielsen PH, Rattei T, Stingl U, Tringe SG, Trojan D, Wentrup C, Woebken D, Pester M, Loy A. 2018. Peatland Acidobacteria with a dissimilatory sulfur metabolism. ISME J 12:1729–1742. doi: 10.1038/s41396-018-0077-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Abicht HK, Mancini S, Karnachuk OV, Solioz M. 2011. Genome sequence of Desulfosporosinus sp. OT, an acidophilic sulfate-reducing bacterium from copper mining waste in Norilsk, Northern Siberia. J Bacteriol 193:6104–6105. doi: 10.1128/JB.06018-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Pester M, Brambilla E, Alazard D, Rattei T, Weinmaier T, Han J, Lucas S, Lapidus A, Cheng J-F, Goodwin L, Pitluck S, Peters L, Ovchinnikova G, Teshima H, Detter JC, Han CS, Tapia R, Land ML, Hauser L, Kyrpides NC, Ivanova NN, Pagani I, Huntmann M, Wei C-L, Davenport KW, Daligault H, Chain PSG, Chen A, Mavromatis K, Markowitz V, Szeto E, Mikhailova N, Pati A, Wagner M, Woyke T, Ollivier B, Klenk H-P, Spring S, Loy A. 2012. Complete genome sequences of Desulfosporosinus orientis DSM765T, Desulfosporosinus youngiae DSM17734T, Desulfosporosinus meridiei DSM13257T, and Desulfosporosinus acidiphilus DSM22704T. J Bacteriol 194:6300–6301. doi: 10.1128/JB.01392-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Abu Laban N, Tan B, Dao A, Foght J. 2015. Draft genome sequence of uncultivated Desulfosporosinus sp. strain Tol-M, obtained by stable isotope probing using [13C6]toluene. Genome Announc 3:e01422-14. doi: 10.1128/genomeA.01422-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Petzsch P, Poehlein A, Johnson DB, Daniel R, Schlömann M, Mühling M. 2015. Genome sequence of the moderately acidophilic sulfate-reducing firmicute Desulfosporosinus acididurans (strain M1T). Genome Announc 3:e00881-15. doi: 10.1128/genomeA.00881-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Mardanov AV, Panova IA, Beletsky AV, Avakyan MR, Kadnikov VV, Antsiferov DV, Banks D, Frank YA, Pimenov NV, Ravin NV, Karnachuk OV. 2016. Genomic insights into a new acidophilic, copper-resistant Desulfosporosinus isolate from the oxidized tailings area of an abandoned gold mine. FEMS Microbiol Ecol 92:fiw111. doi: 10.1093/femsec/fiw111. [DOI] [PubMed] [Google Scholar]
- 35.Stoddard SF, Smith BJ, Hein R, Roller BRK, Schmidt TM. 2015. rrnDB: improved tools for interpreting rRNA gene abundance in bacteria and archaea and a new foundation for future development. Nucleic Acids Res 43:D593–D598. doi: 10.1093/nar/gku1201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Stackebrandt E, Ebers J. 2006. Taxonomic parameters revisited: tarnished gold standards. Microbiol Today 33:152–155. [Google Scholar]
- 37.Varghese NJ, Mukherjee S, Ivanova N, Konstantinidis KT, Mavrommatis K, Kyrpides NC, Pati A. 2015. Microbial species delineation using whole genome sequences. Nucleic Acids Res 43:6761–6771. doi: 10.1093/nar/gkv657. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Rabus R, Venceslau SS, Wöhlbrand L, Voordouw G, Wall JD, Pereira IAC. 2015. A post-genomic view of the ecophysiology, catabolism and biotechnological relevance of sulphate-reducing prokaryotes. Adv Microb Physiol 66:55–321. doi: 10.1016/bs.ampbs.2015.05.002. [DOI] [PubMed] [Google Scholar]
- 39.Huang CJ, Barrett EL. 1991. Sequence analysis and expression of the Salmonella typhimurium asr operon encoding production of hydrogen sulfide from sulfite. J Bacteriol 173:1544–1553. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Weghoff MC, Bertsch J, Müller V. 2015. A novel mode of lactate metabolism in strictly anaerobic bacteria. Environ Microbiol 17:670–677. doi: 10.1111/1462-2920.12493. [DOI] [PubMed] [Google Scholar]
- 41.Zhang Y, Jiang T, Sheng B, Long Y, Gao C, Ma C, Xu P. 2016. Coexistence of two D-lactate-utilizing systems in Pseudomonas putida KT2440. Environ Microbiol Rep 8:699–707. doi: 10.1111/1758-2229.12429. [DOI] [PubMed] [Google Scholar]
- 42.Desguin B, Goffin P, Viaene E, Kleerebezem M, Martin-Diaconescu V, Maroney MJ, Declercq J-P, Soumillion P, Hols P. 2014. Lactate racemase is a nickel-dependent enzyme activated by a widespread maturation system. Nat Commun 5:3615. doi: 10.1038/ncomms4615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Pierce E, Xie G, Barabote RD, Saunders E, Han CS, Detter JC, Richardson P, Brettin TS, Das A, Ljungdahl LG, Ragsdale SW. 2008. The complete genome sequence of Moorella thermoacetica (f. Clostridium thermoaceticum). Environ Microbiol 10:2550–2573. doi: 10.1111/j.1462-2920.2008.01679.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Greening C, Biswas A, Carere CR, Jackson CJ, Taylor MC, Stott MB, Cook GM, Morales SE. 2016. Genomic and metagenomic surveys of hydrogenase distribution indicate H2 is a widely utilised energy source for microbial growth and survival. ISME J 10:761–777. doi: 10.1038/ismej.2015.153. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Schmidt A, Müller N, Schink B, Schleheck D. 2013. A proteomic view at the biochemistry of syntrophic butyrate oxidation in Syntrophomonas wolfei. PLoS One 8:e56905. doi: 10.1371/journal.pone.0056905. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Buckel W, Thauer RK. 2013. Energy conservation via electron bifurcating ferredoxin reduction and proton/Na+ translocating ferredoxin oxidation. Biochim Biophys Acta 1827:94–113. doi: 10.1016/j.bbabio.2012.07.002. [DOI] [PubMed] [Google Scholar]
- 47.Jacks G, Norrström A-C. 2004. Hydrochemistry and hydrology of forest riparian wetlands. For Ecol Manage 196:187–197. doi: 10.1016/j.foreco.2004.01.055. [DOI] [Google Scholar]
- 48.Knorr K-H, Lischeid G, Blodau C. 2009. Dynamics of redox processes in a minerotrophic fen exposed to a water table manipulation. Geoderma 153:379–392. doi: 10.1016/j.geoderma.2009.08.023. [DOI] [Google Scholar]
- 49.Knorr K-H, Blodau C. 2009. Impact of experimental drought and rewetting on redox transformations and methanogenesis in mesocosms of a northern fen soil. Soil Biol Biochem 41:1187–1198. doi: 10.1016/j.soilbio.2009.02.030. [DOI] [Google Scholar]
- 50.Frei S, Knorr K-H, Peiffer S, Fleckenstein JH. 2012. Surface micro-topography causes hot spots of biogeochemical activity in wetland systems: a virtual modeling experiment. J Geophys Res Biogeosci 117:G00N12. [Google Scholar]
- 51.Mojardín L, Salas M. 2016. Global transcriptional analysis of virus-host interactions between phage ϕ29 and Bacillus subtilis. J Virol 90:9293–9304. doi: 10.1128/JVI.01245-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Bergkessel M, Basta DW, Newman DK. 2016. The physiology of growth arrest: uniting molecular and environmental microbiology. Nat Rev Microbiol 14:549–562. doi: 10.1038/nrmicro.2016.107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Teira E, Lekunberri I, Gasol JM, Nieto-Cid M, Alvarez-Salgado XA, Figueiras FG. 2007. Dynamics of the hydrocarbon-degrading Cycloclasticus bacteria during mesocosm-simulated oil spills. Environ Microbiol 9:2551–2562. doi: 10.1111/j.1462-2920.2007.01373.x. [DOI] [PubMed] [Google Scholar]
- 54.Newton RJ, Huse SM, Morrison HG, Peake CS, Sogin ML, McLellan SL. 2013. Shifts in the microbial community composition of Gulf Coast beaches following beach oiling. PLoS One 8:e74265. doi: 10.1371/journal.pone.0074265. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Vergin K, Done B, Carlson C, Giovannoni S. 2013. Spatiotemporal distributions of rare bacterioplankton populations indicate adaptive strategies in the oligotrophic ocean. Aquat Microb Ecol 71:1–13. doi: 10.3354/ame01661. [DOI] [Google Scholar]
- 56.Alazard D, Joseph M, Battaglia-Brunet F, Cayol J-L, Ollivier B. 2010. Desulfosporosinus acidiphilus sp. nov.: a moderately acidophilic sulfate-reducing bacterium isolated from acid mining drainage sediments. Extremophiles 14:305–312. doi: 10.1007/s00792-010-0309-4. [DOI] [PubMed] [Google Scholar]
- 57.Mayeux B, Fardeau M-L, Bartoli-Joseph M, Casalot L, Vinsot A, Labat M. 2013. Desulfosporosinus burensis sp. nov., a spore-forming, mesophilic, sulfate-reducing bacterium isolated from a deep clay environment. Int J Syst Evol Microbiol 63:593–598. doi: 10.1099/ijs.0.035238-0. [DOI] [PubMed] [Google Scholar]
- 58.Sánchez-Andrea I, Stams AJM, Hedrich S, Ňancucheo I, Johnson DB. 2015. Desulfosporosinus acididurans sp. nov.: an acidophilic sulfate-reducing bacterium isolated from acidic sediments. Extremophiles 19:39–47. doi: 10.1007/s00792-014-0701-6. [DOI] [PubMed] [Google Scholar]
- 59.Vandieken V, Niemann H, Engelen B, Cypionka H. 2017. Marinisporobacter balticus gen. nov., sp. nov., Desulfosporosinus nitroreducens sp. nov. and Desulfosporosinus fructosivorans sp. nov., new spore-forming bacteria isolated from subsurface sediments of the Baltic Sea. Int J Syst Evol Microbiol 67:1887–1893. doi: 10.1099/ijsem.0.001883. [DOI] [PubMed] [Google Scholar]
- 60.Schmalenberger A, Drake HL, Küsel K. 2007. High unique diversity of sulfate-reducing prokaryotes characterized in a depth gradient in an acidic fen. Environ Microbiol 9:1317–1328. doi: 10.1111/j.1462-2920.2007.01251.x. [DOI] [PubMed] [Google Scholar]
- 61.Küsel K, Blöthe M, Schulz D, Reiche M, Drake HL. 2008. Microbial reduction of iron and porewater biogeochemistry in acidic peatlands. Biogeosciences 5:1537–1549. doi: 10.5194/bg-5-1537-2008. [DOI] [Google Scholar]
- 62.Overkamp W, Ercan O, Herber M, van Maris AJA, Kleerebezem M, Kuipers OP. 2015. Physiological and cell morphology adaptation of Bacillus subtilis at near-zero specific growth rates: a transcriptome analysis. Environ Microbiol 17:346–363. doi: 10.1111/1462-2920.12676. [DOI] [PubMed] [Google Scholar]
- 63.Goffin P, van de Bunt B, Giovane M, Leveau JHJ, Höppener-Ogawa S, Teusink B, Hugenholtz J. 2010. Understanding the physiology of Lactobacillus plantarum at zero growth. Mol Syst Biol 6:413. doi: 10.1038/msb.2010.67. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Rolfe MD, Rice CJ, Lucchini S, Pin C, Thompson A, Cameron ADS, Alston M, Stringer MF, Betts RP, Baranyi J, Peck MW, Hinton JCD. 2012. Lag phase is a distinct growth phase that prepares bacteria for exponential growth and involves transient metal accumulation. J Bacteriol 194:686–701. doi: 10.1128/JB.06112-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Pirt SJ. 1965. The maintenance energy of bacteria in growing cultures. Proc R Soc Lond B Biol Sci 163:224–231. doi: 10.1098/rspb.1965.0069. [DOI] [PubMed] [Google Scholar]
- 66.Ercan O, Bisschops MMM, Overkamp W, Jørgensen TR, Ram AF, Smid EJ, Pronk JT, Kuipers OP, Daran-Lapujade P, Kleerebezem M. 2015. Physiological and transcriptional responses of different industrial microbes at near-zero specific growth rates. Appl Environ Microbiol 81:5662–5670. doi: 10.1128/AEM.00944-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Tijhuis L, Van Loosdrecht MC, Heijnen JJ. 1993. A thermodynamically based correlation for maintenance gibbs energy requirements in aerobic and anaerobic chemotrophic growth. Biotechnol Bioeng 42:509–519. doi: 10.1002/bit.260420415. [DOI] [PubMed] [Google Scholar]
- 68.Davidson MM, Bisher ME, Pratt LM, Fong J, Southam G, Pfiffner SM, Reches Z, Onstott TC. 2009. Sulfur isotope enrichment during maintenance metabolism in the thermophilic sulfate-reducing bacterium Desulfotomaculum putei. Appl Environ Microbiol 75:5621–5630. doi: 10.1128/AEM.02948-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Kvint K, Nachin L, Diez A, Nyström T. 2003. The bacterial universal stress protein: function and regulation. Curr Opin Microbiol 6:140–145. [DOI] [PubMed] [Google Scholar]
- 70.Jan G, Leverrier P, Pichereau V, Boyaval P. 2001. Changes in protein synthesis and morphology during acid adaptation of Propionibacterium freudenreichii. Appl Environ Microbiol 67:2029–2036. doi: 10.1128/AEM.67.5.2029-2036.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Frees D, Vogensen FK, Ingmer H. 2003. Identification of proteins induced at low pH in Lactococcus lactis. Int J Food Microbiol 87:293–300. [DOI] [PubMed] [Google Scholar]
- 72.Sánchez B, Champomier-Vergès M-C, Collado MDC, Anglade P, Baraige F, Sanz Y, de los Reyes-Gavilán CG, Margolles A, Zagorec M. 2007. Low-pH adaptation and the acid tolerance response of Bifidobacterium longum biotype longum. Appl Environ Microbiol 73:6450–6459. doi: 10.1128/AEM.00886-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Silva J, Carvalho AS, Ferreira R, Vitorino R, Amado F, Domingues P, Teixeira P, Gibbs PA. 2005. Effect of the pH of growth on the survival of Lactobacillus delbrueckii subsp. bulgaricus to stress conditions during spray-drying. J Appl Microbiol 98:775–782. doi: 10.1111/j.1365-2672.2004.02516.x. [DOI] [PubMed] [Google Scholar]
- 74.Lund P, Tramonti A, De Biase D. 2014. Coping with low pH: molecular strategies in neutralophilic bacteria. FEMS Microbiol Rev 38:1091–1125. doi: 10.1111/1574-6976.12076. [DOI] [PubMed] [Google Scholar]
- 75.Albertsen M, Hugenholtz P, Skarshewski A, Nielsen KL, Tyson GW, Nielsen PH. 2013. Genome sequences of rare, uncultured bacteria obtained by differential coverage binning of multiple metagenomes. Nat Biotechnol 31:533–538. doi: 10.1038/nbt.2579. [DOI] [PubMed] [Google Scholar]
- 76.Herbold CW, Lehtovirta-Morley LE, Jung M-Y, Jehmlich N, Hausmann B, Han P, Loy A, Pester M, Sayavedra-Soto LA, Rhee S-K, Prosser JI, Nicol GW, Wagner M, Gubry-Rangin C. 2017. Ammonia-oxidising archaea living at low pH: insights from comparative genomics. Environ Microbiol 19:4939–4952. doi: 10.1111/1462-2920.13971. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Parks DH, Imelfort M, Skennerton CT, Hugenholtz P, Tyson GW. 2015. CheckM: assessing the quality of microbial genomes recovered from isolates, single cells, and metagenomes. Genome Res 25:1043–1055. doi: 10.1101/gr.186072.114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Lartillot N, Lepage T, Blanquart S. 2009. PhyloBayes 3: a Bayesian software package for phylogenetic reconstruction and molecular dating. Bioinformatics 25:2286–2288. doi: 10.1093/bioinformatics/btp368. [DOI] [PubMed] [Google Scholar]
- 79.Quast C, Pruesse E, Yilmaz P, Gerken J, Schweer T, Yarza P, Peplies J, Glöckner FO. 2012. The SILVA ribosomal RNA gene database project: improved data processing and web-based tools. Nucleic Acids Res 41:D590–D596. doi: 10.1093/nar/gks1219. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Pruesse E, Peplies J, Glöckner FO. 2012. SINA: accurate high-throughput multiple sequence alignment of ribosomal RNA genes. Bioinformatics 28:1823–1829. doi: 10.1093/bioinformatics/bts252. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Trifinopoulos J, Nguyen L-T, von Haeseler A, Minh BQ. 2016. W-IQ-TREE: a fast online phylogenetic tool for maximum likelihood analysis. Nucleic Acids Res 44:W232–W235. doi: 10.1093/nar/gkw256. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Notredame C, Higgins DG, Heringa J. 2000. T-Coffee: a novel method for fast and accurate multiple sequence alignment. J Mol Biol 302:205–217. doi: 10.1006/jmbi.2000.4042. [DOI] [PubMed] [Google Scholar]
- 83.Vallenet D, Calteau A, Cruveiller S, Gachet M, Lajus A, Josso A, Mercier J, Renaux A, Rollin J, Rouy Z, Roche D, Scarpelli C, Médigue C. 2017. MicroScope in 2017: an expanding and evolving integrated resource for community expertise of microbial genomes. Nucleic Acids Res 45:D517–D528. doi: 10.1093/nar/gkw1101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.The UniProt Consortium. 2017. UniProt: the universal protein knowledgebase. Nucleic Acids Res 45:D158–D169. doi: 10.1093/nar/gkw1099. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Mitchell A, Chang H-Y, Daugherty L, Fraser M, Hunter S, Lopez R, McAnulla C, McMenamin C, Nuka G, Pesseat S, Sangrador-Vegas A, Scheremetjew M, Rato C, Yong S-Y, Bateman A, Punta M, Attwood TK, Sigrist CJA, Redaschi N, Rivoire C, Xenarios I, Kahn D, Guyot D, Bork P, Letunic I, Gough J, Oates M, Haft D, Huang H, Natale DA, Wu CH, Orengo C, Sillitoe I, Mi H, Thomas PD, Finn RD. 2015. The InterPro protein families database: the classification resource after 15 years. Nucleic Acids Res 43:D213–D221. doi: 10.1093/nar/gku1243. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Jones P, Binns D, Chang H-Y, Fraser M, Li W, McAnulla C, McWilliam H, Maslen J, Mitchell A, Nuka G, Pesseat S, Quinn AF, Sangrador-Vegas A, Scheremetjew M, Yong S-Y, Lopez R, Hunter S. 2014. InterProScan 5: genome-scale protein function classification. Bioinformatics 30:1236–1240. doi: 10.1093/bioinformatics/btu031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Haft DH, Selengut JD, White O. 2003. The TIGRFAMs database of protein families. Nucleic Acids Res 31:371–373. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Overbeek R, Olson R, Pusch GD, Olsen GJ, Davis JJ, Disz T, Edwards RA, Gerdes S, Parrello B, Shukla M, Vonstein V, Wattam AR, Xia F, Stevens R. 2014. The SEED and the Rapid Annotation of microbial genomes using Subsystems Technology (RAST). Nucleic Acids Res 42:D206–D214. doi: 10.1093/nar/gkt1226. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Galperin MY, Makarova KS, Wolf YI, Koonin EV. 2015. Expanded microbial genome coverage and improved protein family annotation in the COG database. Nucleic Acids Res 43:D261–D269. doi: 10.1093/nar/gku1223. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Langmead B, Salzberg SL. 2012. Fast gapped-read alignment with Bowtie 2. Nat Methods 9:357–359. doi: 10.1038/nmeth.1923. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Liao Y, Smyth GK, Shi W. 2014. featureCounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics 30:923–930. doi: 10.1093/bioinformatics/btt656. [DOI] [PubMed] [Google Scholar]
- 92.Love MI, Huber W, Anders S. 2014. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol 15:550. doi: 10.1186/s13059-014-0550-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.R Core Team. 2017. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. [Google Scholar]
- 94.Loy A, Küsel K, Lehner A, Drake HL, Wagner M. 2004. Microarray and functional gene analyses of sulfate-reducing prokaryotes in low-sulfate, acidic fens reveal cooccurrence of recognized genera and novel lineages. Appl Environ Microbiol 70:6998–7009. doi: 10.1128/AEM.70.12.6998-7009.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Katoh K, Standley DM. 2013. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol 30:772–780. doi: 10.1093/molbev/mst010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Huerta-Cepas J, Szklarczyk D, Forslund K, Cook H, Heller D, Walter MC, Rattei T, Mende DR, Sunagawa S, Kuhn M, Jensen LJ, von Mering C, Bork P. 2016. eggNOG 4.5: a hierarchical orthology framework with improved functional annotations for eukaryotic, prokaryotic and viral sequences. Nucleic Acids Res 44:D286–D293. doi: 10.1093/nar/gkv1248. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.