Skip to main content
mSystems logoLink to mSystems
. 2021 Dec 7;6(6):e01042-21. doi: 10.1128/mSystems.01042-21

Omics-Inferred Partitioning and Expression of Diverse Biogeochemical Functions in a Low-O2 Cyanobacterial Mat Community

Sharon L Grim a,*,#, Alexander A Voorhies a,#, Bopaiah A Biddanda b, Sunit Jain a, Stephen C Nold c, Russ Green d,§, Gregory J Dick a,e,
Editor: Hans C Bernsteinf
PMCID: PMC8651085  PMID: 34874776

ABSTRACT

Cyanobacterial mats profoundly influenced Earth’s biological and geochemical evolution and still play important ecological roles in the modern world. However, the biogeochemical functioning of cyanobacterial mats under persistent low-O2 conditions, which dominated their evolutionary history, is not well understood. To investigate how different metabolic and biogeochemical functions are partitioned among community members, we conducted metagenomics and metatranscriptomics on cyanobacterial mats in the low-O2, sulfidic Middle Island sinkhole (MIS) in Lake Huron. Metagenomic assembly and binning yielded 144 draft metagenome assembled genomes, including 61 of medium quality or better, and the dominant cyanobacteria and numerous Proteobacteria involved in sulfur cycling. Strains of a Phormidium autumnale-like cyanobacterium dominated the metagenome and metatranscriptome. Transcripts for the photosynthetic reaction core genes psaA and psbA were abundant in both day and night. Multiple types of psbA genes were expressed from each cyanobacterium, and the dominant psbA transcripts were from an atypical microaerobic type of D1 protein from Phormidium. Further, cyanobacterial transcripts for photosystem I genes were more abundant than those for photosystem II, and two types of Phormidium sulfide quinone reductase were recovered, consistent with anoxygenic photosynthesis via photosystem I in the presence of sulfide. Transcripts indicate active sulfur oxidation and reduction within the cyanobacterial mat, predominately by Gammaproteobacteria and Deltaproteobacteria, respectively. Overall, these genomic and transcriptomic results link specific microbial groups to metabolic processes that underpin primary production and biogeochemical cycling in a low-O2 cyanobacterial mat and suggest mechanisms for tightly coupled cycling of oxygen and sulfur compounds in the mat ecosystem.

IMPORTANCE Cyanobacterial mats are dense communities of microorganisms that contain photosynthetic cyanobacteria along with a host of other bacterial species that play important yet still poorly understood roles in this ecosystem. Although such cyanobacterial mats were critical agents of Earth’s biological and chemical evolution through geological time, little is known about how they function under the low-oxygen conditions that characterized most of their natural history. Here, we performed sequencing of the DNA and RNA of modern cyanobacterial mat communities under low-oxygen and sulfur-rich conditions from the Middle Island sinkhole in Lake Huron. The results reveal the organisms and metabolic pathways that are responsible for both oxygen-producing and non-oxygen-producing photosynthesis as well as interconversions of sulfur that likely shape how much O2 is produced in such ecosystems. These findings indicate tight metabolic reactions between community members that help to explain the limited the amount of O2 produced in cyanobacterial mat ecosystems.

KEYWORDS: cyanobacteria, geomicrobiology, metagenomics, metatranscriptomics, photosynthesis, biogeochemistry, mats, oxygen, sulfur

INTRODUCTION

Cyanobacterial mats host communities of microorganisms that are linked through metabolic interactions in which the products of one metabolism are the substrate for another (14). These metabolic interactions underpinned critical biogeochemical processes throughout Earth’s history (57) and continue to do so in the modern world (2, 4). Cyanobacterial mats have been a prevalent feature of the biosphere for billions of years and strongly influenced the composition of the atmosphere (7, 8). Most prominently, cyanobacteria mediated the oxygenation of Earth’s surface by producing O2 via oxygenic photosynthesis, thus catalyzing a cascade of geobiological transitions that set the stage for complex life (9).

Modern microbial mats have long served as analogs for studying their ancient equivalents, and recent work has made great progress in elucidating cyanobacterial mat processes, organisms, and their interactions (1012). However, relatively little work has been devoted to cyanobacterial mats that inhabit persistently low-O2 and/or sulfidic environments. This is a critical gap in knowledge, because low-O2, sulfidic phototrophic habitats were likely common in the Precambrian (13) and thus prevailed for much of the evolutionary history of cyanobacteria (6, 14). Further, cyanobacteria were likely anoxygenic phototrophs prior to evolving oxygenic photosynthesis (1518), and anoxygenic cyanobacteria may have delayed Earth’s oxygenation during ∼2 billion years of low-O2 conditions in the Proterozoic (6, 19, 20).

Sulfide is a key control of the physiology of cyanobacteria and the biogeochemical cycling of elements in cyanobacterial mats (6). Cyanobacteria typically conduct oxygenic photosynthesis, which is inhibited by sulfide because it blocks photosystem II (PSII) (21). However, some cyanobacteria can tolerate sulfide through a variety of mechanisms, including sulfide-resistant oxygenic photosynthesis, simultaneous operation of oxygenic and anoxygenic photosynthesis, and a complete switch to anoxygenic photosynthesis using sulfide as the electron donor (21). In some strains, sulfide can either inhibit or enhance oxygenic photosynthesis, depending on light availability and sulfide conditions (22). Sulfide-quinone reductase (SQR) is the key enzyme for anoxygenic photosynthesis by cyanobacteria; it oxidizes sulfide and transfers electrons to PSI through the quinone pool, effectively bypassing PSII (2325). SQR is a diverse protein family that has also been linked to sulfide detoxification in cyanobacteria and other phototrophs (6, 24). Although studies have elucidated the physiological responses of cyanobacteria to sulfide and the role of SQR in anoxygenic photosynthesis (21, 26, 27), little is known about transcriptomic controls on cyanobacterial anoxygenic photosynthesis within cyanobacterial mats.

The Middle Island sinkhole (MIS) in Lake Huron, MI, hosts cyanobacterial mats in low-O2, intermittently sulfidic conditions (28). The mats sit atop anoxic, organic-rich sediments in which microbial methanogenesis and sulfate reduction produce methane and sulfide, leading to sharp redox gradients (2932). The mats are metabolically versatile, having the ability to conduct oxygenic photosynthesis, anoxygenic photosynthesis, and chemosynthesis (30, 33, 34). Despite this metabolic versatility, early 16S rRNA gene and metagenomic studies suggested that the mats have low taxonomic diversity, being dominated by just one cyanobacterial genotype, an organism closely related to Phormidium autumnale (29, 30, 34). However, deep 16S rRNA gene sequencing of the mat and underlying sediments revealed a taxonomically diverse microbial community, including numerous groups of sulfate-reducing and sulfur-oxidizing bacteria that are suggested to mediate key biogeochemical processes within and beneath the mat (31). Further, diurnal vertical migration of sulfur-oxidizing bacteria and diatoms exerts a strong influence on the biogeochemistry of the systems and on light availability and photosynthesis in the mats (33, 35).

In order to investigate how different metabolic and geochemical functions are partitioned among community members and expressed over time, we conducted metagenomic analysis on 15 samples collected at seven time points between 2007 and 2012 and metatranscriptomic analysis on six samples taken during day and night in 2012. The Phormidium species was found to dominate transcriptional activity in the MIS mat community and displays gene expression patterns consistent with a mixture of oxygenic and anoxygenic photosynthesis. We also recovered genomes and transcripts of diatoms, sulfate-reducing bacteria, and sulfur-oxidizing bacteria, providing insights into the microbial groups that mediate key biogeochemical processes within the mat.

RESULTS AND DISCUSSION

Environmental setting and conditions.

The environmental and geological setting of the MIS was described previously (28). In May 2012, at the time of collection of samples for metatranscriptomic and metagenomic sequencing (Table S1), the groundwater layer ∼1 m immediately above the mat in the sinkhole had substantially elevated specific conductivity (1,813 μS cm−1 versus 226 μS cm−1 in the ambient lake water), lower and temporally consistent temperature (7 to 9°C), and an average dissolved O2 level of 3.37 mg L−1.

TABLE S1

Summary of samples and accession numbers for the NCBI Short Read Archive (SRA). Download Table S1, PDF file, 0.1 MB (57.1KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

Community composition and function.

Assembly and binning produced 16 high-quality draft metagenome-assembled genomes (MAGs) (>90% completion, <5% redundancy), 45 medium-quality (>50% completion, <10% redundancy), and 79 low-quality draft MAGs (<50% completion, <10% redundancy) according to estimates based on single-copy genes expected to be present (36) (Table S2). In addition, four MAGs had high redundancy (>10%), including three of the most abundant MAGs (Bin_4_1, Bin_1, Bin_235_243; Rhodoferax, Phormidium, and Planktothrix, respectively), which had high coverage and moderate completion (Table S2). For example, the dominant MAG in most samples, Phormidium (Bin_1), had high redundancy (56%) and moderate completeness (70%). Single-copy genes in the Phormidium MAG were on small contigs, consistent with fragmentation of contigs due to high coverage and strain heterogeneity (37, 38), and they were classified taxonomically as various cyanobacteria, as expected based on the lack of available Phormidium genomes (Fig. S1). Thus, these MAGs likely contain contigs from multiple strains of Phormidium.

FIG S1

Contig length and taxonomic classification of single-copy genes (SCG) from the Phormidium bin. Download FIG S1, PDF file, 0.3 MB (291.1KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

TABLE S2

Accession numbers for MAGs that passed NCBI quality filtering. Download Table S2, XLSX file, 0.03 MB (35KB, xlsx) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

While community membership was dynamic across time and space, Phormidium was consistently the dominant organism in the MIS mats (Fig. S2). Other cyanobacteria were also abundant in the mat, including Planktothrix (formerly referred to as Oscillatoria in previous studies of MIS, but its 16S rRNA genes are most similar to those of Planktothrix agardhii and Planktothrix rubescens [39]), Pseudanabaena, and Spirulina. MAGs were also recovered for various Bacteroidetes, Betaproteobacteria, Chloroflexi, Deltaproteobacteria, Epsilonproteobacteria, Firmicutes, Gammaproteobacteria, and Spirochaetes (Table S2). In most cases there were multiple MAGs recovered for each of these taxonomic groups. Many of these groups are commonly found in anoxic or hypoxic sediments (40, 41), and several are enriched in sediments below the mats at the MIS (29, 31).

FIG S2

Average metagenomic (gDNA) and metatranscriptomic (cDNA) coverage of MAGs in each sample. Colors indicate taxonomic designation of MAGs at the genus level, where available. Multiple MAGs may have the same taxonomic designation and thus the same color. Download FIG S2, PDF file, 0.2 MB (175.3KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

To investigate which community members have metabolic pathways for mat biogeochemical processes, we searched the MAGs for key genes involved in carbon metabolism, nitrogen and sulfur cycling, oxygenic and anoxygenic photosynthesis, and other energy metabolisms. Cyanobacteria were the dominant phototrophs in terms of genomic abundance; Phormidium had a mean genomic coverage of over 200×, though based on high redundancy (56%), multiple strains are present (Table S2). While two putative diatom MAGs (Bin_3_1 and Bin_3_3) had low average coverages (0.31 and 6.19×), their chloroplasts were very abundant (up to 230× coverage), likely reflecting their high copy number per cell and easier assembly than the nuclear genome (Fig. S3). Marker genes of anoxygenic photosynthesis, including photosynthetic reactions center (pufM and pufL) and bacteriochlorophyll synthesis (bchB and bchL), were present in Rhodoferax (Bin_4_1) and Chloroflexi (Bin_120) MAGs (Table S2). Diatoms and Chloroflexi are often associated with cyanobacterial mats; migratory diatoms play important roles in nitrogen cycling in MIS mats and sediments (35), and Chloroflexi engage in tight metabolic interactions with cyanobacteria (2, 3, 10, 42).

FIG S3

Average metagenomic (gDNA) and metatranscriptomic (cDNA) coverage of putative diatom chloroplasts. Binned scaffolds in putative diatoms were identified as chloroplasts based on the presence of photosynthetic reaction center genes (psbA). Download FIG S3, PDF file, 0.1 MB (62.2KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

The MAGs of various proteobacteria revealed organisms involved in sulfur cycling. Key genes for dissimilatory sulfite reductase (dsrA) involved in sulfate reduction were present in deltaproteobacterial genomic bins, including one unclassified Desulfobacteraceae member, one unclassifiable Desulfobulbaceae member, and one Desulfobacula member (Table S2). Based on their relatively high abundance, we infer that these sulfate-reducing bacteria were present directly in the cyanobacterial mat (43, 44) rather than the alternative that the sequences could be due to contamination of the mat by underlying sediments. Related sulfate-reducing bacteria are associated with anoxygenic bacteria in Lake Mahoney (45), with cyanobacterial mats at Guerrero Negro (46, 47), and in nearby mats of chemolithotrophic sulfur oxidizers influenced by the same groundwater as MIS (48).

Potential for oxidation of elemental sulfur using reverse dissimilatory sulfite reductase (rdsrA) was detected in genomic bins of Arcobacter (Epsilonproteobacteria), several Betaproteobacteria, Thiothrix, and Thioploca and in unbinned scaffolds putatively belonging to Beggiatoa (Table S2). Thiothrix, Thioploca, and Beggiatoa are likely the white filamentous bacteria observed directly underneath the cyanobacterial mat (29) and in a nearby artesian fountain fed by the same groundwater (48). They can migrate on diel cycles and influence the balance of oxygenic versus anoxygenic photosynthesis by modulating light available to phototrophs when covering the mat (33). These large sulfur-oxidizing bacteria likely contribute to the substantial rates of chemosynthesis measured previously (30) and likely influence cyanobacterial photosynthesis by consuming sulfide. Potential for thiosulfate oxidation, indicated by the presence of soxA, was observed in betaproteobacterial, deltaproteobacterial, and gammaproteobacterial bins (Table S2). Finally, the mmoC gene, for methane oxidation, was identified within a MAG classified as Methylococcales (Table S2).

Whole-community transcriptomics.

Metatranscriptomic sequencing was conducted to investigate the in situ metabolic activity of the MIS mat community members. Although transcript abundance is not directly proportional to protein abundance or enzymatic activity, transcriptomics provides valuable insights into which community members and metabolic pathways are active at the time of sampling and their response to environmental conditions (49). In order to evaluate the influence of light availability on gene expression in MIS mats, three samples collected in 2012 at 1 p.m. and three collected at 1 a.m. were studied. In terms of relative abundance, transcripts mapped to MAGs from Phormidium, Bacteroidetes, Thiotricaceae, and the putative diatom dominated the metatranscriptome (Fig. S2; Table S2). Other significant contributors (>1× mean coverage) to the transcript pool were bins from Paludibacter and other Bacteroidetes members, Rhodoferax (Betaproteobacteria), Chloroflexi, Planktothrix, and a variety of unidentified bins (Table S2). Mapping of metatranscriptomic data to marker genes and MAGs provided a picture of the organisms responsible for metabolic/biogeochemical processes within the mat in day and night (Fig. 1; Table S2).

FIG 1.

FIG 1

Detection of transcripts from marker genes of key metabolic/biogeochemical processes in MAGs (see Table S2 for details). Symbols are colored according to the time of day at which transcripts were detected: white, day; black, night; gray, both night and day. “X” indicates that the gene was observed in the MAG, but no transcripts were detected.

Transcripts involved in phototrophy.

We next focused our transcriptomic analysis on key genes for photosynthesis. Core components of the reaction centers of PSI and PSII, encoded by psaA and psbA genes, respectively, are degraded at an enhanced rate compared to other proteins due to absorption of excess light energy from photosynthesis (50, 51). This leads to higher cellular demand for protein and likely explains the high abundance of transcripts we observed for these genes. The most abundant transcripts for psaA and psbA genes were from Phormidium and the diatom, with minor contributions from Planktothrix, Spirulina, and Pseudanabaena (Fig. 2). Included in our analyses were multiple versions of the cyanobacterial psbA genes, encoding the D1 subunit of PSII, which are expressed according to light and redox conditions (52, 53) and have been suggested to be involved in sulfide tolerance and/or anoxygenic photosynthesis in cyanobacteria (54). Phormidium contained three of the four psbA types, and type 3 had the most transcripts (Fig. 2). This type of psbA is expressed during microaerobic and/or dynamic redox conditions (55, 56). The other cyanobacterium that was abundant in metagenomic data sets, Planktothrix, had type 3 and 4 psbA genes, with similar relative abundance of transcripts in day and night but at much lower levels than Phormidium.

FIG 2.

FIG 2

Relative abundance of transcripts for photosystem genes normalized by total number of sequences in each sample. Log-transformed transcript abundance in the day (white) and night (gray) of genes encoding photosystem I (psaA) and photosystem II (psbA) is shown for each MAG (top) with box-and-whisker plots. Boxes represent the 25th to 75th percentiles, the inside line indicates the median, and whiskers extend to minimum and maximum values. Observations are overlaid as points. The x-axis labels “psbA_2,” “psbA_3,” and “psbA_4” refer to psbA types (see the text).

Twenty-one of 32 psaA and psbA genes had transcripts that were more abundant at night than in the day, although these differences were not statistically significant (Fig. 2). When normalized by the number of transcript reads recruited to each bin, which removes effects of transcriptomic variability across the whole community on transcript counts for genes within each bin, 9 of 21 genes had transcripts that were more abundant at night (Fig. S4). These patterns contrast those in several laboratory studies of cyanobacterial transcription, which found highest expression of photosynthesis reaction core genes during the day (5762). Several field studies have also shown highest expression of PSII genes during the day (6365). One possible explanation is that whereas many previous studies focused on oxygenic unicellular cyanobacteria that typically undergo rapid cell division (59, 66), Phormidium species are filamentous and typically much slower growing (0.07 to 0.5 day−1, depending on light and nutrient availability) (67, 68). There is a precedent for high transcription of photosynthetic genes in the dark and lower transcription in the light; in Synechococcus sp. strain PCC 7002, transcript levels of several psbA genes were constant across several conditions, including light and dark (69).

FIG S4

Relative abundance of transcripts from psbA and psaA genes in day and night, normalized to the number of transcripts recruited to each MAG. Log-transformed bin-specific TPM of transcript abundance in day (white) and night (grey) of genes encoding photosystem I (psaA) and photosystem II (psbA) are shown for each MAG (top). The x-axis labels “psbA_2”, “psbA_3” and “psbA_4” refer to psbA types (see the text). Boxes represent the 25th to 75th percentiles, the inside line indicates the median, and whiskers extend to minimum and maximum values. Observations are overlaid as points. Download FIG S4, PDF file, 0.3 MB (340.6KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

Phototrophic genes were among the most highly expressed genes in marine surface waters collected 3 h before sunrise (70). Photosystem I genes in thrombolites were constitutively expressed, with nearly even transcript abundance at midday and midnight (64). Anoxygenic phototrophs such as Chloroflexi and Chlorobi also express structural components of the photosynthetic apparatus at night (11, 71). Finally, our results could also be explained in part by decreased rates of afternoon photosynthesis that have been observed in cyanobacterial mats, including sharp drops at midday (7274), which have been attributed to limitation of dissolved inorganic carbon (2). It should also be noted that transcript levels do not necessarily reflect protein abundance; under some circumstances, psbA messages accumulate without synthesis of the D1:2 protein (75, 76).

The abundance of transcripts obtained from Phormidium PSII genes indicates that genes for oxygenic photosynthesis were transcriptionally active at the time of sampling. However, Phormidium had on average more than two times higher abundance of transcripts for PSI genes than PSII genes (psbA-to-psaA ratio < 0.5) (Fig. 3). In contrast, the diatom chloroplast recruited more than 25 times more transcripts to PSII genes (psbA) than PSI genes (psaA). However, likely due to high variability of the abundance of diatom transcripts for these genes, this large difference in the ratio of transcript abundance from PSII and PSI genes was not statistically significant.

FIG 3.

FIG 3

Ratio of transcript abundances for photosystem II (psbA) to photosystem I (psaA) genes in day and night, shown for each MAG (top) with box-and-whisker plots. Boxes represent the 25th to 75th percentiles, the inside line indicates the median, and whiskers extend to minimum and maximum values. Observations are overlaid as points.

Other MIS cyanobacteria exhibited an intermediate ratio, with 2 to 7 times more PSII than PSI transcripts. A high ratio of PSII to PSI transcripts was also found in Prochlorococcus, a unicellular marine planktonic cyanobacterium (59). We infer that the higher relative abundance of PSI transcripts in Phormidium (and perhaps other MIS cyanobacteria) reflects transcriptional regulation, either via downregulation of PSII genes or upregulation of PSI genes, to conduct anoxygenic photosynthesis in the presence of sulfide. Although to our knowledge these are the first transcriptional data from anoxygenic cyanobacteria, they are consistent with the physiological shift toward PSII-independent anoxygenic photosynthesis that was reported previously (21, 27), with a decrease in the stoichiometry of PSII-PSI in response to sulfide (77), and with genes for anoxygenic photosynthesis being inducible via transcriptional regulation (24, 78, 79). The stoichiometry of PSII-PSI can also be regulated according to light levels (80); the PSII/PSI ratio is lower at lower light levels, which also favors anoxygenic photosynthesis in the MIS mat system (33).

Sulfide quinone oxidoreductase (SQR) transfers electrons from sulfide to PSI during anoxygenic photosynthesis (24). Of the five cyanobacterial sqr homologs recovered from the MIS community, the Phormidium SQR had the highest abundance of transcripts, from both its type I and II sqr genes (Fig. 4). The bin-normalized transcripts-per-million (TPM) value of Phormidium’s type II SQR was significantly higher than that for SQRI (P < 0.05). While MAGs of Planktothrix and Pseudanabaena have SQRs (6), transcripts for these genes were not observed. The Phormidium SQRs showed transcript abundance comparable to that of the PSI genes psaL and psaX (Fig. S5). Little is known about how anoxygenic photosynthesis and sulfide tolerance are regulated at the genetic level in cyanobacteria. Expression of type II SQR for sulfide detoxification in Synechocystis sp. strain PCC6803 (79) and type I SQR for anoxygenic photosynthesis in Geitlerinema sp. strain 9228 (24, 81, 82) is inducible by sulfide. Both constitutive expression (25, 83, 84) and inducible expression (85, 86) of sqr have been observed in anoxygenic bacteria. There was little metatranscriptomic evidence of anoxygenic photosynthesis by anoxygenic bacteria (i.e., Chloroflexi or Betaproteobacteria). Genes for photosynthetic reaction cores (pufM and pufL) and bacteriochlorophyll (bchB and bchL) were not highly expressed, with 0 or 1 read mapped in all samples. Overall, these results suggest that the cyanobacteria are largely responsible for anoxygenic photosynthesis previously measured in MIS mats (30, 33).

FIG 4.

FIG 4

Relative abundance of transcripts for sulfide quinone reductase (SQR) type I (putatively involved in anoxygenic photosynthesis) and type II (putatively involved in sulfide detoxification) genes from the Phormidium MAG, normalized by transcripts recruited to the MAG for each sample. Boxes represent the 25th to 75th percentiles, the inside line indicates the median, and whiskers extend to minimum and maximum values. Observations are overlaid as points.

FIG S5

Relative abundance of transcripts from sqr, psaL, and psaX from the Phormidium bin in day and night, normalized to the number of sequences retrieved from each sample. Sample-normalized TPM of transcript abundance in the day (white) and night (grey) are shown as box-and-whisker plots in which boxes represent the 25th to 75th percentiles, the inside line indicates the median, and whiskers extend to minimum and maximum values. Observations are overlaid as points. Download FIG S5, PDF file, 0.2 MB (164.1KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

The most highly expressed terminal oxidase for respiration in Phormidium was a cytochrome bd-type oxidase (Fig. S6), which has exceptionally high affinity for O2, with a Km of 3 to 8 nM (87). The high transcriptional activity of this low-O2 respiratory oxidase is consistent with adaptation to low-O2 conditions for extended time periods.

FIG S6

Relative abundance of transcripts from genes encoding terminal respiratory oxidases in the Phormidium MAG, normalized to the number of transcripts recruited to the Phormidium MAG in each sample. Log-transformed bin-specific TPM of transcript abundance in the day (white) and night (grey) of genes are shown as box-and-whisker plots in which boxes represent the 25th to 75th percentiles, the inside line indicates the median, and whiskers extend to minimum and maximum values. Observations are overlaid as points. Download FIG S6, PDF file, 0.1 MB (154KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

Transcripts involved in sulfur cycling and carbon fixation.

Transcripts of seven different dsrA genes were observed, and the presence of dsrD on the same scaffold (Fig. S7) was used to confirm inclusion of these genes in the dissimilatory sulfite reductase pathway (dsr genes). dsrD is useful a marker of sulfite reduction because it is absent from organisms that use homologous rdsrA genes for sulfur oxidation (88, 89). Transcriptionally active reductive dsrA genes were present in seven MAGs representing six genera of Deltaproteobacteria, with most transcripts coming from Desulfococcus (Deltaproteobacteria; Desulfobacterales), followed by Desulfomicrobium (Fig. 5). These results reveal the organisms responsible for active sulfate reduction within the cyanobacterial mat, which has been measured at high rates by 35SO42− tracer studies (32). For most sulfate-reducing MAGs, dsrA transcripts were detected and even more abundant during the day, suggesting sulfate reduction during the photosynthetic period and likely metabolic interactions with cyanobacteria via the cycling of sulfur and/or carbon (6, 90). These sulfate-reducing bacteria are also present in sediments underlying the cyanobacterial mat (31).

FIG 5.

FIG 5

Relative abundance of transcripts for key genes for dissimilatory sulfite reduction (dsrA) and reverse dissimilatory sulfite reduction (rdsrA) normalized by total number of sequences in each sample. Boxes represent the 25th to 75th percentiles, the inside line indicates the median, and whiskers extend to minimum and maximum values. Observations are overlaid as points.

FIG S7

Transcriptional relative abundance and MAG bin statistics of genes involved in dissimilatory sulfite reduction (dsr; dsrA and dsrD genes) and reverse dsr (rdsrA and dsrEFH). In the first and third panels, the values represent sample-normalized TPM, shown as box-and-whisker plots in which boxes represent the 25th to 75th percentiles, the inside line indicates the median, and whiskers extend to minimum and maximum values. Observations are overlaid as points. The second and fourth panels present the percent completion of MAG bins, when available, in which the genes are located. Many genes were unable to be binned into MAG bins (NA), despite being transcriptionally active in day (white) or night (grey). The average coverage in metatranscriptomic samples is presented in Table S2. Download FIG S7, PDF file, 0.1 MB (85.9KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

Seven rdsrA genes for sulfur oxidation were observed, including those in MAGs from three Betaproteobacteria (two unclassified and one member of the Hydrogenophilales), one classified only as proteobacteria, and one Thiobacillus organism (Fig. 5). Two unbinned genes most similar to Thiothrix and Beggiatoa (Gammaproteobacteria) were also recovered. With the exception of the unbinned proteobacterial gene, all of these rdsrA genes had more transcripts at night. Transcripts from soxA genes for thiosulfate oxidation were detected, with those from Rhodoferax (Betaproteobacteria) and unbinned representatives of the Campylobacterales (Epsilonproteobacteria) having the highest abundance of transcripts (Fig. S8). Transcripts of genes for sulfide oxidation via flavocytochrome c sulfide dehydrogenase (fcc) were also observed in bins from the Betaproteobacteria and Gammaproteobacteria, though their sample-normalized transcript abundance was nearly an order of magnitude lower than those of dsrA.

FIG S8

Abundance of transcripts from genes encoding oxidation of various sulfur species. The sample-normalized TPM of day (white) and night (grey) transcripts of the genes fccAB (sulfide oxidation using flavocytochrome c sulfide dehydrogenase) and soxA (thiosulfate oxidation) are plotted by taxonomy of the MAG (if available) or scaffold (if unbinned). Boxes represent the 25th to 75th percentiles, the inside line indicates the median, and whiskers extend to minimum and maximum values. Observations are overlaid as points. Only MAGs or scaffolds that recruited transcripts are shown here; 40 additional soxA genes and 40 additional fccAB genes were identified in the metagenomic dataset but were not represented in the metatranscriptome. Download FIG S8, PDF file, 0.1 MB (69.8KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

To assess sources of primary production at MIS, we measured abundance of transcripts encoding key genes of four autotrophic pathways: ribulose-1,5-bisphosphate carboxylase/oxygenase (RuBisCO) for the Calvin cycle (rbcL), ATP citrate lyase (aclB) for the reverse tricarboxylic acid cycle, CO dehydrogenase/acetyl coenzyme A (acetyl-CoA) synthase (acsB) for the Wood-Ljungdahl pathway, and malyl-CoA/(S)-citramalyl-CoA lyase (mcl), malonyl-CoA reductase/3-hydroxypropionate dehydrogenase, and 3-hydroxypropionyl-CoA dehydratase (mcr) for the 3-hydroxypropionate cycle (Fig. 6). RuBisCO had the highest transcript abundance of any autotrophic pathway. Phormidium, Planktothrix, and a diatom chloroplast actively transcribed rbcL, and the cyanobacteria especially were active at night. Thiotricaceae also expressed rbcL but at substantially lower levels than the phototrophs. Several Desulfobacterales (Deltaproteobacteria) MAGs expressed the Wood-Ljungdahl genes, with higher expression during the day. We did not observe transcriptional activity of aclB, mcl, or mcr genes in any samples.

FIG 6.

FIG 6

Relative abundance of transcripts for key genes in active autotrophic pathways normalized by total number of sequences in each sample. The sample-normalized TPM of each gene in daytime (white) and nighttime (gray) are plotted as box and whiskers in which boxes represent the 25th to 75th percentiles, the inside line indicates the median, and whiskers extend to minimum and maximum values. Observations are overlaid as points. Bin and taxonomy of indicator genes for RuBisCO (rbcL) and the Wood-Ljungdahl pathway (acsB) are indicated on the x axis.

Conclusions.

Our data on taxon-specific genomic content and gene expression patterns provide insights into the microbial players and pathways that mediate biogeochemistry in anoxygenic cyanobacterial mats, which likely underpinned critical aspects of Earth’s geobiological evolution but have been understudied in the modern world. The metatranscriptomic data revealed that a Phormidium autumnale-like cyanobacterium previously found to be the dominant community member (29, 30) is also responsible for the majority of transcripts for photosynthesis, underscoring its essential role in the MIS ecosystem. Taken together, low expression ratios of PSII genes to PSI genes and expression of SQR indicate that Phormidium within MIS mats conducts anoxygenic photosynthesis with sulfide as the electron donor, consistent with previous geochemical measurements at the MIS (30, 33) as well as studies of other members of the Oscillatoriales (24, 91). Hence, the Phormidium population appears to be metabolically versatile, capable of both oxygenic and anoxygenic photosynthesis. It is not clear whether this phototrophic versatility stems from niche adaptation among closely related ecotypes (e.g., differential activity of strains that are oxygenic and anoxygenic specialists) or true cellular versatility in which Phormidium cells switch pathways depending on sulfide concentration. Further, our bulk sampling of mats was not sensitive to the vertical microgradients of sulfide concentration, so we are unable to evaluate potential vertical stratification of oxygenic/anoxygenic photosynthesis within the mat.

This study provides a picture of how metabolisms encoding specific biogeochemical functions are partitioned among mat community members (Fig. 1). Primary production occurs via oxygenic photosynthesis by cyanobacteria and diatoms, anoxygenic photosynthesis by cyanobacteria (with minor contributions potentially from Chloroflexi and Betaproteobacteria), and chemosynthesis via sulfur oxidation by Thiotrichales and Chromatiales (Gammaproteobacteria) and Epsilonproteobacteria. Sulfide sets the stage for these metabolisms and is produced via sulfate and/or sulfur reduction by several different groups of Deltaproteobacteria both within the cyanobacterial mat and in underlying sediments. Bacteroidetes are the major heterotrophs, consuming organic carbon released as photosynthetic exudate or via viral lysis (92). Although the spatial arrangement of these processes within the mat remains unresolved, sulfide, oxygen, and organic carbon are likely rapidly cycled between the organisms producing and consuming them (14). Such tightly coupled interactions would help explain why there is little mat-derived carbon sequestered in the sediments (93). Tight coupling of O2 production and consumption metabolisms, together with substantial primary production by anoxygenic photosynthesis, also helps explain the limited net O2 production by the cyanobacterial mat when measured in bulk (30). Overall, these findings emphasize the importance of microbial metabolic interactions in shaping biogeochemical processes in cyanobacterial mats under low-O2 conditions, which dominated the long evolutionary history of cyanobacteria and played key but poorly understood roles in Earth’s major geobiological turning points.

MATERIALS AND METHODS

Sample collection and sequencing.

This study used samples and metagenomic and metatranscriptomic sequence data produced as described by Voorhies et al. (92). Fifteen mat samples were collected by scuba divers from the R/V Storm between 2007 and 2012 from within a 100-m area of the Middle Island sinkhole (45.1984°N, 83.32721°W) by hand push cores of sediments, mat, and overlying groundwater (Table S1). Cores were rapidly transferred to the surface, and mats were separated from underlying sediments and submerged in RNAlater immediately shipboard. Less than 5 min elapsed between collection and preservation. Before preservation, mat samples were quickly washed with groundwater to remove as much sediment as possible. Mat structure ranged from conical structures we refer to as “fingers” (30) to prostrate mat. In May 2012, at the time of metagenomic and metatranscriptomic sequencing, conductivity, temperature, and dissolved O2 in the overlying lake water as well as the groundwater above the mats were measured by a YSI 6600 multiparameter sonde.

DNA was extracted and processed for shotgun metagenomic sequencing as previously described (30). Samples were sequenced using an Illumina Hi Seq 2000 (paired end, 100 bp) instrument at the University of Michigan DNA Sequencing Core. In 2012, three samples of mat were collected at approximately 1 p.m. (day) and 1 a.m. (night) from within a 9-m2 sampling area. RNA was extracted from these six samples, randomly amplified with the MessageAmp II-Bacteria kit (Ambion), and converted to cDNA using the SuperScript double-stranded-cDNA synthesis kit (Invitrogen), as previously described (70). In the interest of cost efficiency and to minimize sample handling, rRNA was not removed (94, 95). cDNA was sequenced at the University of Michigan DNA Sequencing Core on an Illumina Hi Seq 2000 instrument producing paired-end reads.

Assembly and genomic analysis.

In order to optimize assembly of genomes from low abundance members, a total of 922 million sequence reads from all 15 genomic DNA samples were combined and coassembled and binned by two different methods. We used IDBA-UD (96) for assembly and checked the results against previous assemblies of MIS mats that used multiple sequencing platforms and assembly programs, including a previously published metagenome based on 454 (28, 30), and assemblies of Illumina data from individual samples performed with Velvet (97). Specifically, the assembly was checked by verifying recovery of key genes and MAGs that were observed in the previous assemblies and by manual curation using the Integrated Genome Viewer (98) and Geneious (99) to visualize reads mapped to contigs and genes by BWA (Burrows-Wheeler aligner) (100) and look for signs of misassembly (e.g., discontinuities in coverage). Multiple strategies and software were used to generate metagenome-assembled-genome (MAG) bins. The first strategy used CONCOCT (101) to automate binning by differential coverage and tetranucleotide frequency for the subset of contigs that were 5 kb and larger. The resulting bins were refined manually in anvi’o (102, 103) and assigned taxonomy via Centrifuge (104) and CheckM (105). Likely due to high coverage and putative strain heterogeneity (37, 38), the 12.5-Mbp MAG bin representing the dominant cyanobacterium Phormidium had high completion (95.0%) but poor contamination (97.1%) metrics (106) (Table S2). For this bin as well as 6 other cyanobacterial MAG bins from the initial refinement, 6,011 contigs that were previously unbinned due to their short length (1,000 bp to 4,999 bp) were assigned to bins on the basis of similar coverage, nucleotide composition (tetranucleotide frequencies), and taxonomy via manual refinement in anvi’o.

We also employed a second, purely automated binning strategy for comparison. EukRep (107) removed eukaryotic contigs from the data set, and MetaBAT (108) used differential coverage and tetranucleotide frequency to generate MAGs from contigs 1,500 bp and longer. We again used CheckM to taxonomically identify MAG bins and tracked 16S rRNA, psbA, and sqr genes from the previously extracted cyanobacterial bins to identify their counterparts in the MetaBAT bins. Contigs previously assigned to Phormidium were poorly binned in this method. dRep (109) was used to pick the best representative bin from the two methods. Though 10 of the bins from CONCOCT+anvi’o were retained through dRep, the MetaBAT-generated bins were more often picked because they generally had lower estimates of contamination and strain heterogeneity. Putative Phormidium scaffolds that were not binned by MetaBAT were manually evaluated in anvi’o, and retained as the representative Phormidium bin in this analysis. Gene calling and functional annotation was performed by the Joint Genome Institute’s Integrated Microbial Genomes Expert Review portal (https://img.jgi.doe.gov/cgi-bin/mer/main.cgi) (110).

Coverage of contigs by cDNA and DNA sequence reads from each sample was assessed by mapping reads to contigs using BWA (100) with default settings. Raw counts of cDNA reads (referred to here as counts) for each predicted protein-coding gene were determined using the IMG-derived coordinates of gene start and stop sites, along with the mapping information. rRNA genes erroneously called as protein-coding genes were identified by BLASTn against the SILVA SSU and LSU database, release 123 (111), and removed. The python script HTSeq.scripts.count from HTSeq (112) extracted transcript counts that unambiguously mapped to genes. In targeted searches for metabolic genes of interest, we identified 20 partial “genes” that were not suitable lengths when IMG-determined start and stop sites were used and were of appropriate lengths when partial “genes” were incorporated immediately upstream or downstream. Thus, for metatranscriptomic analyses, the counts of these partial genes were merged.

For analysis of metatranscriptomic data, only genes with at least two counts were considered. Two different normalization methods were used to analyze the metatranscriptomic data, depending on the question. First, transcript abundance was normalized by total mRNA reads recovered in each sample to calculate relative abundance of transcripts at the gene level. This metric is a function of both the organism abundance and expression per gene copy and provides a measure of total contribution to the transcript pool for each gene. Second, to compare relative gene expression within genomic bins (and remove the effect of dynamic community-wide transcript and organism abundance), we normalized relative abundance of transcripts by number of mRNAs mapped to each genome bin. To account for variability in sequencing effort between samples and for the impact of gene and read lengths, gene expression levels were normalized using TPM (113). To evaluate differences in expression levels between organisms of assembled metagenomic bins, TPM for bin-specific genes were also calculated with the denominator consisting of only reads recruited to the bin of interest. Statistical testing was conducted on the sample- and bin-normalized TPM of genes in RStudio using Kruskal-Wallis nonparametric tests and paired t tests, corrected with a Benjamini-Hochberg false discovery rate (q) of 0.05.

Data availability.

Sequences from this study are available from NCBI under BioProject no. PRJNA72255. Reads from all 15 metagenomes and 6 metatranscriptomes are available in NCBI’s Sequence Read Archive (Table S1). Accession numbers for MAGs that passed NCBI quality filtering are provided in Table S2.

Supplementary Material

Reviewer comments
reviewer-comments.pdf (701.8KB, pdf)

ACKNOWLEDGMENTS

Divers and crew of the R/V Storm NOAA and the Thunder Bay National Marine Sanctuary provided critical dive support and assistance with sampling and ship time. We thank the University of Michigan DNA sequencing core for DNA sequencing.

This work was supported by NSF grants EAR1035955 and EAR 1637066 to G.J.D., NSF grant EAR1035957 to B.A.B., the University of Michigan CCMB Pilot Grant to G.J.D., and the Scott Turner Award to A.A.V.

We declare no conflicts of interest.

Contributor Information

Gregory J. Dick, Email: gdick@umich.edu.

Hans C. Bernstein, UiT—The Arctic University of Norway

REFERENCES

  • 1.Paerl HW, Pinckney JL, Steppe TF. 2000. Cyanobacterial-bacterial mat consortia: examining the functional unit of microbial survival and growth in extreme environments. Environ Microbiol 2:11–26. doi: 10.1046/j.1462-2920.2000.00071.x. [DOI] [PubMed] [Google Scholar]
  • 2.Stal LJ. 2012. Cyanobacterial mats and stromatolites, p 65–125. In Whitton BA (ed), Ecology of cyanobacteria II: their diversity in space and time. Springer, Heidelberg, Germany. [Google Scholar]
  • 3.Nold S, Ward D. 1996. Photosynthate partitioning and fermentation in hot spring microbial mat communities. Appl Environ Microbiol 62:4598–4607. doi: 10.1128/aem.62.12.4598-4607.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Stal LJ, Bolhuis H, Cretoiu MS. 2019. Phototrophic marine benthic microbiomes: the ecophysiology of these biological entities. Environ Microbiol 21:1529–1551. doi: 10.1111/1462-2920.14494. [DOI] [PubMed] [Google Scholar]
  • 5.Ward LM, Kirschvink JL, Fischer WW. 2016. Timescales of oxygenation following the evolution of oxygenic photosynthesis. Orig Life Evol Biosph 46:51–65. doi: 10.1007/s11084-015-9460-3. [DOI] [PubMed] [Google Scholar]
  • 6.Dick GJ, Grim SL, Klatt JM. 2018. Controls on O2 production in cyanobacterial mats and implications for Earth's oxygenation. Annu Rev Earth Planet Sci 46:123–147. doi: 10.1146/annurev-earth-082517-010035. [DOI] [Google Scholar]
  • 7.Sanchez-Baracaldo P, Bianchini G, Wilson JD, Knoll AH. 2021. Cyanobacteria and biogeochemical cycles through Earth history. Trends Microbiol doi: 10.1016/j.tim.2021.05.008. [DOI] [PubMed] [Google Scholar]
  • 8.Hoehler TM, Bebout BM, Des Marais DJ. 2001. The role of microbial mats in the production of reduced gases on the early Earth. Nature 412:324–327. doi: 10.1038/35085554. [DOI] [PubMed] [Google Scholar]
  • 9.Falkowski PG, Isozaki Y. 2008. The story of O2. Science 322:540–542. doi: 10.1126/science.1162641. [DOI] [PubMed] [Google Scholar]
  • 10.Burow LC, Woebken D, Marshall IPG, Lindquist EA, Bebout BM, Prufert-Bebout L, Hoehler TM, Tringe SG, Pett-Ridge J, Weber PK, Spormann AM, Singer SW. 2013. Anoxic carbon flux in photosynthetic microbial mats as revealed by metatranscriptomics. ISME J 7:817–829. doi: 10.1038/ismej.2012.150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Klatt CG, Liu ZF, Ludwig M, Kuhl M, Jensen SI, Bryant DA, Ward DM. 2013. Temporal metatranscriptomic patterning in phototrophic Chloroflexi inhabiting a microbial mat in a geothermal spring. ISME J 7:1775–1789. doi: 10.1038/ismej.2013.52. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Harris JK, Caporaso JG, Walker JJ, Spear JR, Gold NJ, Robertson CE, Hugenholtz P, Goodrich J, McDonald D, Knights D, Marshall P, Tufo H, Knight R, Pace NR. 2013. Phylogenetic stratigraphy in the Guerrero Negro hypersaline microbial mat. ISME J 7:50–60. doi: 10.1038/ismej.2012.79. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Hanson TE, Luther GW, Findlay AJ, MacDonald DJ, Hess D. 2013. Phototrophic sulfide oxidation: environmental insights and a method for kinetic analysis. Front Microbiol 4:382. doi: 10.3389/fmicb.2013.00382. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Hamilton TL, Bryant DA, Macalady JL. 2016. The role of biology in planetary evolution: cyanobacterial primary production in low-oxygen Proterozoic oceans. Environ Microbiol 18:325–340. doi: 10.1111/1462-2920.13118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Olson JM. 2006. Photosynthesis in the Archean Era. Photosynth Res 88:109–117. doi: 10.1007/s11120-006-9040-5. [DOI] [PubMed] [Google Scholar]
  • 16.Blankenship R, Sadekar S, Raymond J. 2007. The evolutionary transition from anoxygenic to oxygenic photosynthesis, p 21–35. In Falkowski PG, Knoll AH (ed), Evolution of primary producers in the sea. Elsevier, Amsterdam, The Netherlands. [Google Scholar]
  • 17.Fischer WW, Hemp J, Johnson JE. 2016. Evolution of oxygenic photosynthesis. Annu Rev Earth Planet Sci 44:647–683. doi: 10.1146/annurev-earth-060313-054810. [DOI] [Google Scholar]
  • 18.Hamilton TL. 2019. The trouble with oxygen: the ecophysiology of extant phototrophs and implications for the evolution of oxygenic photosynthesis. Free Radic Biol Med 140:233–249. doi: 10.1016/j.freeradbiomed.2019.05.003. [DOI] [PubMed] [Google Scholar]
  • 19.Johnston DT, Wolfe-Simon F, Pearson A, Knoll AH. 2009. Anoxygenic photosynthesis modulated Proterozoic oxygen and sustained Earth's middle age. Proc Natl Acad Sci USA 106:16925–16929. doi: 10.1073/pnas.0909248106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Ozaki K, Thompson KJ, Simister RL, Crowe SA, Reinhard CT. 2019. Anoxygenic photosynthesis and the delayed oxygenation of Earth's atmosphere. Nat Commun 10:3026. doi: 10.1038/s41467-019-10872-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Cohen Y, Jørgensen BB, Revsbech NP, Poplawski R. 1986. Adaptation to hydrogen sulfide of oxygenic and anoxygenic photosynthesis among cyanobacteria. Appl Environ Microbiol 51:398–407. doi: 10.1128/aem.51.2.398-407.1986. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Klatt JM, Haas S, Yilmaz P, Beer D, Polerecky L. 2015. Hydrogen sulfide can inhibit and enhance oxygenic photosynthesis in a cyanobacterium from sulfidic springs. Environ Microbiol 17:3301–3313. doi: 10.1111/1462-2920.12791. [DOI] [PubMed] [Google Scholar]
  • 23.Arieli B, Shahak Y, Taglicht D, Hauska G, Padan E. 1994. Purification and characterization of sulfide-quinone reductase, a novel enzyme driving anoxygenic photosynthesis in Oscillatoria-Limnetica. J Biol Chem 269:5705–5711. doi: 10.1016/S0021-9258(17)37518-X. [DOI] [PubMed] [Google Scholar]
  • 24.Bronstein M, Schutz M, Hauska G, Padan E, Shahak Y. 2000. Cyanobacterial sulfide-quinone reductase: cloning and heterologous expression. J Bacteriol 182:3336–3344. doi: 10.1128/JB.182.12.3336-3344.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Schütz M, Shahak Y, Padan E, Hauska G. 1997. Sulfide-quinone reductase from Rhodobacter capsulatus. Purification, cloning, and expression. J Biol Chem 272:9890–9894. doi: 10.1074/jbc.272.15.9890. [DOI] [PubMed] [Google Scholar]
  • 26.Miller SR, Bebout BM. 2004. Variation in sulfide tolerance of photosystem II in phylogenetically diverse cyanobacteria from sulfidic habitats. Appl Environ Microbiol 70:736–744. doi: 10.1128/AEM.70.2.736-744.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Jørgensen BB, Cohen Y, Revsbech NP. 1986. Transition from anoxygenic to oxygenic photosynthesis in a Microcoleus chthonoplastes cyanobacterial mat. Appl Environ Microbiol 51:408–417. doi: 10.1128/aem.51.2.408-417.1986. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Ruberg SA, Kendall ST, Biddanda BA, Black T, Nold SC, Lusardi WR, Green R, Casserley T, Smith E, Sanders TG, Lang GA, Constant SA. 2008. Observations of the Middle Island sinkhole in Lake Huron—a unique hydrogeologic and glacial creation of 400 million years. Mar Technol Soc J 42:12–21. doi: 10.4031/002533208787157633. [DOI] [Google Scholar]
  • 29.Nold SC, Pangborn JB, Zajack HA, Kendall ST, Rediske RR, Biddanda BA. 2010. Benthic bacterial diversity in submerged sinkhole ecosystems. Appl Environ Microbiol 76:347–351. doi: 10.1128/AEM.01186-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Voorhies AA, Biddanda B, Kendall ST, Jain S, Marcus DN, Nold SC, Sheldon ND, Dick GJ. 2012. Cyanobacterial life at low O2: community genomics and function reveal metabolic versatility and extremely low diversity in a Great Lakes sinkhole mat. Geobiology 10:250–267. doi: 10.1111/j.1472-4669.2012.00322.x. [DOI] [PubMed] [Google Scholar]
  • 31.Kinsman-Costello L, Sheik CS, Sheldon ND, Burton GA, Costello D, Marcus DN, Den Uyl PA, Dick GJ. 2017. Groundwater shapes sediment biogeochemistry and microbial diversity in a submerged Great Lake sinkhole. Geobiology 15:225–239. doi: 10.1111/gbi.12215. [DOI] [PubMed] [Google Scholar]
  • 32.Gomes ML, Klatt JM, Dick GJ, Grim SL, Rico KI, Medina MJ, Ziebis W, Kinsman-Costello LE, Sheldon ND, Fike DA. 2021. Sedimentary pyrite sulfur isotope compositions preserve signatures of the surface microbial mat environment in sediments underlying low-oxygen cyanobacterial mats. Geobiology doi: 10.1111/gbi.12466. [DOI] [PubMed] [Google Scholar]
  • 33.Klatt JM, Chennu A, Arbic BK, Biddanda BA, Dick GJ. 2021. Possible link between Earth’s rotation rate and oxygenation. Nat Geosci 14:564–570. doi: 10.1038/s41561-021-00784-3. [DOI] [Google Scholar]
  • 34.Snider MJ, Biddanda BA, Lindback M, Grim SL, Dick GJ. 2017. Versatile photophysiology of compositionally similar cyanobacterial mat communities inhabiting submerged sinkholes of Lake Huron. Aquat Microb Ecol 79:63–78. doi: 10.3354/ame01813. [DOI] [Google Scholar]
  • 35.Merz E, Dick GJ, de Beer D, Grim S, Hubener T, Littmann S, Olsen K, Stuart D, Lavik G, Marchant HK, Klatt JM. 2021. Nitrate respiration and diel migration patterns of diatoms are linked in sediments underneath a microbial mat. Environ Microbiol 23:1422–1435. doi: 10.1111/1462-2920.15345. [DOI] [PubMed] [Google Scholar]
  • 36.Bowers RM, Kyrpides NC, Stepanauskas R, Harmon-Smith M, Doud D, Reddy TBK, Schulz F, Jarett J, Rivers AR, Eloe-Fadrosh EA, Tringe SG, Ivanova NN, Copeland A, Clum A, Becraft ED, Malmstrom RR, Birren B, Podar M, Bork P, Weinstock GM, Garrity GM, Dodsworth JA, Yooseph S, Sutton G, Glockner FO, Gilbert JA, Nelson WC, Hallam SJ, Jungbluth SP, Ettema TJG, Tighe S, Konstantinidis KT, Liu WT, Baker BJ, Rattei T, Eisen JA, Hedlund B, McMahon KD, Fierer N, Knight R, Finn R, Cochrane G, Karsch-Mizrachi I, Tyson GW, Rinke C, Genome Standards C, Lapidus A, Meyer F, Yilmaz P, Parks DH, Genome Standards Consortium , et al. 2017. Minimum information about a single amplified genome (MISAG) and a metagenome-assembled genome (MIMAG) of bacteria and archaea. Nat Biotechnol 35:725–731. doi: 10.1038/nbt.3893. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Handley KM, Bartels D, O'Loughlin EJ, Williams KH, Trimble WL, Skinner K, Gilbert JA, Desai N, Glass EM, Paczian T, Wilke A, Antonopoulos D, Kemner KM, Meyer F. 2014. The complete genome sequence for putative H(2)- and S-oxidizer Candidatus Sulfuricurvum sp., assembled de novo from an aquifer-derived metagenome. Environ Microbiol 16:3443–3462. doi: 10.1111/1462-2920.12453. [DOI] [PubMed] [Google Scholar]
  • 38.Hug LA, Thomas BC, Sharon I, Brown CT, Sharma R, Hettich RL, Wilkins MJ, Williams KH, Singh A, Banfield JF. 2016. Critical biogeochemical functions in the subsurface are associated with bacteria from new phyla and little studied lineages. Environ Microbiol 18:159–173. doi: 10.1111/1462-2920.12930. [DOI] [PubMed] [Google Scholar]
  • 39.Suda S, Watanabe MM, Otsuka S, Mahakahant A, Yongmanitchai W, Nopartnaraporn N, Liu Y, Day JG. 2002. Taxonomic revision of water-bloom-forming species of oscillatorioid cyanobacteria. Int J Syst Evol Microbiol 52:1577–1595. doi: 10.1099/00207713-52-5-1577. [DOI] [PubMed] [Google Scholar]
  • 40.Tsertova N, Kisand A, Tammert H, Kisand V. 2011. Low seasonal variability in community composition of sediment bacteria in large and shallow lake. Environ Microbiol Rep 3:270–277. doi: 10.1111/j.1758-2229.2010.00221.x. [DOI] [PubMed] [Google Scholar]
  • 41.Yau S, Lauro FM, Williams TJ, DeMaere MZ, Brown MV, Rich J, Gibson JAE, Cavicchioli R. 2013. Metagenomic insights into strategies of carbon conservation and unusual sulfur biogeochemistry in a hypersaline Antarctic lake. ISME J 7:1944–1961. doi: 10.1038/ismej.2013.69. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.van der Meer MT, Klatt CG, Wood J, Bryant DA, Bateson MM, Lammerts L, Schouten S, Damste JS, Madigan MT, Ward DM. 2010. Cultivation and genomic, nutritional, and lipid biomarker characterization of Roseiflexus strains closely related to predominant in situ populations inhabiting Yellowstone hot spring microbial mats. J Bacteriol 192:3033–3042. doi: 10.1128/JB.01610-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Jonkers HM, Ludwig R, Wit R, Pringault O, Muyzer G, Niemann H, Finke N, Beer D. 2003. Structural and functional analysis of a microbial mat ecosystem from a unique permanent hypersaline inland lake: 'La Salada de Chiprana' (NE Spain). FEMS Microbiol Ecol 44:175–189. doi: 10.1016/S0168-6496(02)00464-6. [DOI] [PubMed] [Google Scholar]
  • 44.Canfield DE, Des Marais DJ. 1991. Aerobic sulfate reduction in microbial mats. Science 251:1471–1473. doi: 10.1126/science.11538266. [DOI] [PubMed] [Google Scholar]
  • 45.Hamilton TL, Bovee RJ, Thiel V, Sattin SR, Mohr W, Schaperdoth I, Vogl K, Gilhooly WP, III, Lyons TW, Tomsho LP, Schuster SC, Overmann J, Bryant DA, Pearson A, Macalady JL. 2014. Coupled reductive and oxidative sulfur cycling in the phototrophic plate of a meromictic lake. Geobiology 12:451–468. doi: 10.1111/gbi.12092. [DOI] [PubMed] [Google Scholar]
  • 46.Fike DA, Gammon CL, Ziebis W, Orphan VJ. 2008. Micron-scale mapping of sulfur cycling across the oxycline of a cyanobacterial mat: a paired nanoSIMS and CARD-FISH approach. ISME J 2:749–759. doi: 10.1038/ismej.2008.39. [DOI] [PubMed] [Google Scholar]
  • 47.Fike DA, Finke N, Zha J, Blake G, Hoehler TM, Orphan VJ. 2009. The effect of sulfate concentration on (sub)millimeter-scale sulfide d34S in hypersaline cyanobacterial mats over the diurnal cycle. Geochim Cosmochim Acta 73:6187–6204. doi: 10.1016/j.gca.2009.07.006. [DOI] [Google Scholar]
  • 48.Sharrar AM, Flood BE, Bailey JV, Jones DS, Biddanda BA, Ruberg SA, Marcus DN, Dick GJ. 2017. Novel large sulfur bacteria in the metagenomes of groundwater-fed chemosynthetic microbial mats in the Lake Huron basin. Front Microbiol 8:791. doi: 10.3389/fmicb.2017.00791. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Moran MA, Satinsky B, Gifford SM, Luo H, Rivers A, Chan LK, Meng J, Durham BP, Shen C, Varaljay VA, Smith CB, Yager PL, Hopkinson BM. 2013. Sizing up metatranscriptomics. ISME J 7:237–243. doi: 10.1038/ismej.2012.94. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Mattoo AK, Hoffman-Falk H, Marder JB, Edelman M. 1984. Regulation of protein-metabolism—coupling of photosynthetic electron-transport to in-vivo degradation of the rapidly metabolized 32-kilodalton protein of the chloroplast membranes. Proc Natl Acad Sci USA 81:1380–1384. doi: 10.1073/pnas.81.5.1380. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Soitamo AJ, Zhou G, Clarke AK, Oquist G, Gustafsson P, Aro EM. 1996. Over-production of the D1:2 protein makes Synechococcus cells more tolerant to photoinhibition of photosystem II. Plant Mol Biol 30:467–478. doi: 10.1007/BF00049325. [DOI] [PubMed] [Google Scholar]
  • 52.Cardona T. 2015. A fresh look at the evolution and diversification of photochemical reaction centers. Photosynth Res 126:111–134. doi: 10.1007/s11120-014-0065-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Cardona T, Murray JW, Rutherford WA. 2015. Origin and evolution of water oxidation before the last common ancestor of the cyanobacteria. Mol Biol Evol 32:1310–1328. doi: 10.1093/molbev/msv024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Grim SL, Dick GJ. 2016. Photosynthetic versatility in the genome of Geitlerinema sp. PCC 9228 (formerly Oscillatoria limnetica ‘Solar Lake’), a model anoxygenic photosynthetic cyanobacterium. Front Microbiol 7:1546. doi: 10.3389/fmicb.2016.01546. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Mulo P, Sicora C, Aro EM. 2009. Cyanobacterial psbA gene family: optimization of oxygenic photosynthesis. Cell Mol Life Sci 66:3697–3710. doi: 10.1007/s00018-009-0103-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Sicora CI, Ho FM, Salminen T, Styring S, Aro EM. 2009. Transcription of a “silent” cyanobacterial psbA gene is induced by microaerobic conditions. Biochim Biophys Acta 1787:105–112. doi: 10.1016/j.bbabio.2008.12.002. [DOI] [PubMed] [Google Scholar]
  • 57.Shi T, Ilikchyan I, Rabouille S, Zehr JP. 2010. Genome-wide analysis of diel gene expression in the unicellular N-2-fixing cyanobacterium Crocosphaera watsonii WH 8501. ISME J 4:621–632. doi: 10.1038/ismej.2009.148. [DOI] [PubMed] [Google Scholar]
  • 58.Stockel J, Jacobs JM, Elvitigala TR, Liberton M, Welsh EA, Polpitiya AD, Gritsenko MA, Nicora CD, Koppenaal DW, Smith RD, Pakrasi HB. 2011. Diurnal rhythms result in significant changes in the cellular protein complement in the cyanobacterium Cyanothece 51142. PLoS One 6:e16680. doi: 10.1371/journal.pone.0016680. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Zinser ER, Lindell D, Johnson ZI, Futschik ME, Steglich C, Coleman ML, Wright MA, Rector T, Steen R, McNulty N, Thompson LR, Chisholm SW. 2009. Choreography of the transcriptome, photophysiology, and cell cycle of a minimal photoautotroph, Prochlorococcus. PLoS One 4:e5135. doi: 10.1371/journal.pone.0005135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Labiosa RG, Arrigo KR, Tu CJ, Bhaya D, Bay S, Grossman AR, Shrager J. 2006. Examination of diel changes in global transcript accumulation in Synechocystis (cyanobacteria). J Phycol 42:622–636. doi: 10.1111/j.1529-8817.2006.00217.x. [DOI] [Google Scholar]
  • 61.Colon-Lopez MS, Sherman LA. 1998. Transcriptional and translational regulation of photosystem I and II genes in light-dark- and continuous-light-grown cultures of the unicellular cyanobacterium Cyanothece sp. strain ATCC 51142. J Bacteriol 180:519–526. doi: 10.1128/JB.180.3.519-526.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Straub C, Quillardet P, Vergalli J, de Marsac NT, Humbert JF. 2011. A day in the life of Microcystis aeruginosa strain PCC 7806 as revealed by a transcriptomic analysis. PLoS One 6:e16208. doi: 10.1371/journal.pone.0016208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Jensen SI, Steunou AS, Bhaya D, Kuhl M, Grossman AR. 2011. In situ dynamics of O2, pH and cyanobacterial transcripts associated with CCM, photosynthesis and detoxification of ROS. ISME J 5:317–328. doi: 10.1038/ismej.2010.131. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Louyakis AS, Gourle H, Casaburi G, Bonjawo RME, Duscher AA, Foster JS. 2018. A year in the life of a thrombolite: comparative metatranscriptomics reveals dynamic metabolic changes over diel and seasonal cycles. Environ Microbiol 20:842–861. doi: 10.1111/1462-2920.14029. [DOI] [PubMed] [Google Scholar]
  • 65.Hornlein C, Confurius-Guns V, Stal LJ, Bolhuis H. 2018. Daily rhythmicity in coastal microbial mats. NPJ Biofilms Microbiomes 4:11. doi: 10.1038/s41522-018-0054-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Vijayan V, Zuzow R, O'Shea EK. 2009. Oscillations in supercoiling drive circadian gene expression in cyanobacteria. Proc Natl Acad Sci USA 106:22564–22568. doi: 10.1073/pnas.0912673106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Litchman E. 2000. Growth rates of phytoplankton under fluctuating light. Freshw Biol 44:223–235. doi: 10.1046/j.1365-2427.2000.00559.x. [DOI] [Google Scholar]
  • 68.Litchman E, Steiner D, Bossard P. 2003. Photosynthetic and growth responses of three freshwater algae to phosphorus limitation and daylength. Freshw Biol 48:2141–2148. doi: 10.1046/j.1365-2427.2003.01157.x. [DOI] [Google Scholar]
  • 69.Ludwig M, Bryant DA. 2011. Transcription profiling of the model cyanobacterium Synechococcus sp. strain PCC 7002 by Next-Gen (SOLiD™) sequencing of cDNA. Front Microbiol 2:41. doi: 10.3389/fmicb.2011.00041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Frias-Lopez J, Shi Y, Tyson GW, Coleman ML, Schuster SC, Chisholm SW, DeLong EF. 2008. Microbial community gene expression in ocean surface waters. Proc Natl Acad Sci USA 105:3805–3810. doi: 10.1073/pnas.0708897105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Liu ZF, Klatt CG, Wood JM, Rusch DB, Ludwig M, Wittekindt N, Tomsho LP, Schuster SC, Ward DM, Bryant DA. 2011. Metatranscriptomic analyses of chlorophototrophs of a hot-spring microbial mat. ISME J 5:1279–1290. doi: 10.1038/ismej.2011.37. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Villbrandt M, Stal LJ, Krumbein WE. 1990. Interactions between nitrogen-fixation and oxygenic photosynthesis in a marine cyanobacterial mat. FEMS Microbiol Ecol 74:59–72. doi: 10.1111/j.1574-6968.1990.tb04052.x. [DOI] [Google Scholar]
  • 73.Paerl HW, Bebout BM, Prufert LE. 1989. Naturally occurring patterns of oxygenic photosynthesis and N2 fixation in a marine microbial mat: physiological and ecological ramifications, p 326–341. In Cohen Y, Rosenberg E (ed), Microbial mats: physiological ecology of benthic communities. American Society for Microbiology, Washington, DC. [Google Scholar]
  • 74.Storch TA, Saunders GW, Ostrofsky ML. 1990. Diel nitrogen fixation by cyanobacterial surface blooms in Sanctuary Lake, Pennsylvania. Appl Environ Microbiol 56:466–471. doi: 10.1128/aem.56.2.466-471.1990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Campbell D, Clarke AK, Gustafsson P, Oquist G. 1999. Oxygen-dependent electron flow influences photosystem II function and psbA gene expression in the cyanobacterium Synechococcus sp PCC 7942. Physiol Plant 105:746–755. doi: 10.1034/j.1399-3054.1999.105420.x. [DOI] [Google Scholar]
  • 76.Sippola K, Aro EM. 2000. Expression of psbA genes is regulated at multiple levels in the cyanobacterium Synechococcus sp PCC 7942. Photochem Photobiol 71:706–714. doi:. [DOI] [PubMed] [Google Scholar]
  • 77.Garcia-Pichel F, Castenholz RW. 1990. Comparative anoxygenic photosynthetic capacity in 7 strains of a thermophilic cyanobacterium. Arch Microbiol 153:344–351. doi: 10.1007/BF00249003. [DOI] [Google Scholar]
  • 78.Shahak Y, Arieli B, Binder B, Padan E. 1987. Sulfide-dependent photosynthetic electron flow coupled to proton translocation in thylakoids of the cyanobacterium Oscillatoria limnetica. Arch Biochem Biophys 259:605–615. doi: 10.1016/0003-9861(87)90527-3. [DOI] [PubMed] [Google Scholar]
  • 79.Nagy CI, Vass I, Rakhely G, Vass IZ, Toth A, Duzs A, Peca L, Kruk J, Kos P. 2014. Coregulated genes link sulfide:quinone oxidoreductase and arsenic metabolism in Synechocysis sp. strain PCC6803. J Bacteriol 196:3430–3440. doi: 10.1128/JB.01864-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Fujita Y, Murakami A, Aizawa K, Ohki K. 1994. Short-term and long-term adaption of the photosynthetic apparatus: homeostatic properties of thylakoids, p 677–692. In Bryant DA (ed), The molecular biology of Cyanobacteria. Kluwer Academic Publishers, Dordrecht, The Netherlands. [Google Scholar]
  • 81.Arieli B, Binder B, Shahak Y, Padan E. 1989. Sulfide induction of synthesis of a periplasmic protein in the cyanobacterium Oscillatoria limnetica. J Bacteriol 171:699–702. doi: 10.1128/jb.171.2.699-702.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Arieli B, Padan E, Shahak Y. 1991. Sulfide-induced sulfide-quinone reductase activity in thylakoids of Oscillatoria limnetica. J Biol Chem 266:104–111. doi: 10.1016/S0021-9258(18)52408-X. [DOI] [PubMed] [Google Scholar]
  • 83.Reinartz M, Tschape J, Bruser T, Truper HG, Dahl C. 1998. Sulfide oxidation in the phototrophic sulfur bacterium Chromatium vinosum. Arch Microbiol 170:59–68. doi: 10.1007/s002030050615. [DOI] [PubMed] [Google Scholar]
  • 84.Shahak Y, Arieli B, Padan E, Hauska G. 1992. Sulfide quinone reductase (SQR) activity in Chlorobium. FEBS Lett 299:127–130. doi: 10.1016/0014-5793(92)80230-e. [DOI] [PubMed] [Google Scholar]
  • 85.Griesbeck C, Hauska G, Schutz M. 2000. Biological sulfide oxidation: sulfide-quinone reductase (SQR), the primary reaction. Recent Res Dev Microbiol 4:179–203. [Google Scholar]
  • 86.Chan LK, Morgan-Kiss RM, Hanson TE. 2009. Functional analysis of three sulfide:quinone oxidoreductase homologs in Chlorobaculum tepidum. J Bacteriol 191:1026–1034. doi: 10.1128/JB.01154-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Morris RL, Schmidt TM. 2013. Shallow breathing: bacterial life at low O(2). Nat Rev Microbiol 11:205–212. doi: 10.1038/nrmicro2970. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Anantharaman K, Hausmann B, Jungbluth SP, Kantor RS, Lavy A, Warren LA, Rappe MS, Pester M, Loy A, Thomas BC, Banfield JF. 2018. Expanded diversity of microbial groups that shape the dissimilatory sulfur cycle. ISME J 12:1715–1728. doi: 10.1038/s41396-018-0078-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Rabus R, Venceslau SS, Wohlbrand L, Voordouw G, Wall JD, Pereira IA. 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]
  • 90.Klatt JM, Gomez-Saez GV, Meyer S, Ristova PP, Yilmaz P, Granitsiotis MS, Macalady JL, Lavik G, Polerecky L, Buhring SI. 2020. Versatile cyanobacteria control the timing and extent of sulfide production in a Proterozoic analog microbial mat. ISME J 14:3024–3037. doi: 10.1038/s41396-020-0734-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Cohen Y, Padan E, Shilo M. 1975. Facultative anoxygenic photosynthesis in cyanobacterium oscillatoria-limnetica. J Bacteriol 123:855–861. doi: 10.1128/jb.123.3.855-861.1975. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Voorhies AA, Eisenlord SD, Marcus DN, Duhaime MB, Biddanda BA, Cavalcoli JD, Dick GJ. 2016. Ecological and genetic interactions between cyanobacteria and viruses in a low-O2 mat community inferred through metagenomics and metatranscriptomics. Environ Microbiol 18:358–371. doi: 10.1111/1462-2920.12756. [DOI] [PubMed] [Google Scholar]
  • 93.Nold SC, Bellecourt MJ, Kendall ST, Ruberg SA, Sanders TG, Klump JV, Biddanda BA. 2013. Underwater sinkhole sediments sequester Lake Huron's carbon. Biogeochemistry 115:235–250. doi: 10.1007/s10533-013-9830-8. [DOI] [Google Scholar]
  • 94.Stewart FJ. 2013. Preparation of microbial community cDNA for metatranscriptomic analysis in marine plankton. Methods Enzymol 531:187–218. doi: 10.1016/B978-0-12-407863-5.00010-1. [DOI] [PubMed] [Google Scholar]
  • 95.Dick GJ. 2018. Genomic approaches in Earth and environmental sciences. Wiley Blackwell, Hoboken, NJ. [Google Scholar]
  • 96.Peng Y, Leung HCM, Yiu SM, Chin FYL. 2012. IDBA-UD: a de novo assembler for single-cell and metagenomic sequencing data with highly uneven depth. Bioinformatics 28:1420–1428. doi: 10.1093/bioinformatics/bts174. [DOI] [PubMed] [Google Scholar]
  • 97.Zerbino DR, Birney E. 2008. Velvet: algorithms for de novo short read assembly using de Bruijn graphs. Genome Res 18:821–829. doi: 10.1101/gr.074492.107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Robinson JT, Thorvaldsdottir H, Winckler W, Guttman M, Lander ES, Getz G, Mesirov JP. 2011. Integrative genomics viewer. Nat Biotechnol 29:24–26. doi: 10.1038/nbt.1754. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Biomatters. 2013. Geneious, v6.1.5. www.geneious.com.
  • 100.Li H, Durbin R. 2009. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics 25:1754–1760. doi: 10.1093/bioinformatics/btp324. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Alneberg J, Bjarnason BS, de Bruijn I, Schirmer M, Quick J, Ijaz UZ, Lahti L, Loman NJ, Andersson AF, Quince C. 2014. Binning metagenomic contigs by coverage and composition. Nat Methods 11:1144–1146. doi: 10.1038/nmeth.3103. [DOI] [PubMed] [Google Scholar]
  • 102.Eren AM, Esen OC, Quince C, Vineis JH, Morrison HG, Sogin ML, Delmont TO. 2015. Anvi'o: an advanced analysis and visualization platform for 'omics data. PeerJ 3:e1319. doi: 10.7717/peerj.1319. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Eren AM, Kiefl E, Shaiber A, Veseli I, Miller SE, Schechter MS, Fink I, Pan JN, Yousef M, Fogarty EC, Trigodet F, Watson AR, Esen OC, Moore RM, Clayssen Q, Lee MD, Kivenson V, Graham ED, Merrill BD, Karkman A, Blankenberg D, Eppley JM, Sjodin A, Scott JJ, Vazquez-Campos X, McKay LJ, McDaniel EA, Stevens SLR, Anderson RE, Fuessel J, Fernandez-Guerra A, Maignien L, Delmont TO, Willis AD. 2021. Community-led, integrated, reproducible multi-omics with anvi'o. Nat Microbiol 6:3–6. doi: 10.1038/s41564-020-00834-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Kim D, Song L, Breitwieser FP, Salzberg SL. 2016. Centrifuge: rapid and sensitive classification of metagenomic sequences. Genome Res 26:1721–1729. doi: 10.1101/gr.210641.116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.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]
  • 106.Campbell JH, O’Donoghue P, Campbell AG, Schwientek P, Sczyrba A, Woyke T, Söll D, Podar M. 2013. UGA is an additional glycine codon in uncultured SR1 bacteria from the human microbiota. Proc Natl Acad Sci U S A 110:5540–5545. doi: 10.1073/pnas.1303090110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.West PT, Probst AJ, Grigoriev IV, Thomas BC, Banfield JF. 2018. Genome-reconstruction for eukaryotes from complex natural microbial communities. Genome Res 28:569–580. doi: 10.1101/gr.228429.117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Kang DD, Froula J, Egan R, Wang Z. 2015. MetaBAT, an efficient tool for accurately reconstructing single genomes from complex microbial communities. PeerJ 3:e1165. doi: 10.7717/peerj.1165. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Olm MR, Brown CT, Brooks B, Banfield JF. 2017. dRep: a tool for fast and accurate genomic comparisons that enables improved genome recovery from metagenomes through de-replication. ISME J 11:2864–2868. doi: 10.1038/ismej.2017.126. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Markowitz VM, Mavromatis K, Ivanova NN, Chen IM, Chu K, Kyrpides NC. 2009. IMG ER: a system for microbial genome annotation expert review and curation. Bioinformatics 25:2271–2278. doi: 10.1093/bioinformatics/btp393. [DOI] [PubMed] [Google Scholar]
  • 111.Quast C, Pruesse E, Yilmaz P, Gerken J, Schweer T, Yarza P, Peplies J, Glockner FO. 2013. 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]
  • 112.Anders S, Pyl PT, Huber W. 2015. HTSeq–a Python framework to work with high-throughput sequencing data. Bioinformatics 31:166–169. doi: 10.1093/bioinformatics/btu638. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Wagner GP, Kin K, Lynch VJ. 2012. Measurement of mRNA abundance using RNA-seq data: RPKM measure is inconsistent among samples. Theory Biosci 131:281–285. doi: 10.1007/s12064-012-0162-3. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

TABLE S1

Summary of samples and accession numbers for the NCBI Short Read Archive (SRA). Download Table S1, PDF file, 0.1 MB (57.1KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

FIG S1

Contig length and taxonomic classification of single-copy genes (SCG) from the Phormidium bin. Download FIG S1, PDF file, 0.3 MB (291.1KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

TABLE S2

Accession numbers for MAGs that passed NCBI quality filtering. Download Table S2, XLSX file, 0.03 MB (35KB, xlsx) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

FIG S2

Average metagenomic (gDNA) and metatranscriptomic (cDNA) coverage of MAGs in each sample. Colors indicate taxonomic designation of MAGs at the genus level, where available. Multiple MAGs may have the same taxonomic designation and thus the same color. Download FIG S2, PDF file, 0.2 MB (175.3KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

FIG S3

Average metagenomic (gDNA) and metatranscriptomic (cDNA) coverage of putative diatom chloroplasts. Binned scaffolds in putative diatoms were identified as chloroplasts based on the presence of photosynthetic reaction center genes (psbA). Download FIG S3, PDF file, 0.1 MB (62.2KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

FIG S4

Relative abundance of transcripts from psbA and psaA genes in day and night, normalized to the number of transcripts recruited to each MAG. Log-transformed bin-specific TPM of transcript abundance in day (white) and night (grey) of genes encoding photosystem I (psaA) and photosystem II (psbA) are shown for each MAG (top). The x-axis labels “psbA_2”, “psbA_3” and “psbA_4” refer to psbA types (see the text). Boxes represent the 25th to 75th percentiles, the inside line indicates the median, and whiskers extend to minimum and maximum values. Observations are overlaid as points. Download FIG S4, PDF file, 0.3 MB (340.6KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

FIG S5

Relative abundance of transcripts from sqr, psaL, and psaX from the Phormidium bin in day and night, normalized to the number of sequences retrieved from each sample. Sample-normalized TPM of transcript abundance in the day (white) and night (grey) are shown as box-and-whisker plots in which boxes represent the 25th to 75th percentiles, the inside line indicates the median, and whiskers extend to minimum and maximum values. Observations are overlaid as points. Download FIG S5, PDF file, 0.2 MB (164.1KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

FIG S6

Relative abundance of transcripts from genes encoding terminal respiratory oxidases in the Phormidium MAG, normalized to the number of transcripts recruited to the Phormidium MAG in each sample. Log-transformed bin-specific TPM of transcript abundance in the day (white) and night (grey) of genes are shown as box-and-whisker plots in which boxes represent the 25th to 75th percentiles, the inside line indicates the median, and whiskers extend to minimum and maximum values. Observations are overlaid as points. Download FIG S6, PDF file, 0.1 MB (154KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

FIG S7

Transcriptional relative abundance and MAG bin statistics of genes involved in dissimilatory sulfite reduction (dsr; dsrA and dsrD genes) and reverse dsr (rdsrA and dsrEFH). In the first and third panels, the values represent sample-normalized TPM, shown as box-and-whisker plots in which boxes represent the 25th to 75th percentiles, the inside line indicates the median, and whiskers extend to minimum and maximum values. Observations are overlaid as points. The second and fourth panels present the percent completion of MAG bins, when available, in which the genes are located. Many genes were unable to be binned into MAG bins (NA), despite being transcriptionally active in day (white) or night (grey). The average coverage in metatranscriptomic samples is presented in Table S2. Download FIG S7, PDF file, 0.1 MB (85.9KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

FIG S8

Abundance of transcripts from genes encoding oxidation of various sulfur species. The sample-normalized TPM of day (white) and night (grey) transcripts of the genes fccAB (sulfide oxidation using flavocytochrome c sulfide dehydrogenase) and soxA (thiosulfate oxidation) are plotted by taxonomy of the MAG (if available) or scaffold (if unbinned). Boxes represent the 25th to 75th percentiles, the inside line indicates the median, and whiskers extend to minimum and maximum values. Observations are overlaid as points. Only MAGs or scaffolds that recruited transcripts are shown here; 40 additional soxA genes and 40 additional fccAB genes were identified in the metagenomic dataset but were not represented in the metatranscriptome. Download FIG S8, PDF file, 0.1 MB (69.8KB, pdf) .

Copyright © 2021 Grim et al.

This content is distributed under the terms of the Creative Commons Attribution 4.0 International license.

Reviewer comments
reviewer-comments.pdf (701.8KB, pdf)

Data Availability Statement

Sequences from this study are available from NCBI under BioProject no. PRJNA72255. Reads from all 15 metagenomes and 6 metatranscriptomes are available in NCBI’s Sequence Read Archive (Table S1). Accession numbers for MAGs that passed NCBI quality filtering are provided in Table S2.


Articles from mSystems are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES