<<

Geobiology (2014), 12, 451–468 DOI: 10.1111/gbi.12092

Coupled reductive and oxidative cycling in the phototrophic plate of a T. L. HAMILTON,1 R. J. BOVEE,2 V. THIEL,3 S. R. SATTIN,2 W. MOHR,2 I. SCHAPERDOTH,1 K. VOGL,3 W. P. GILHOOLY III,4 T. W. LYONS,5 L. P. TOMSHO,3 S. C. SCHUSTER,3,6 J. OVERMANN,7 D. A. BRYANT,3,6,8 A. PEARSON2 AND J. L. MACALADY1 1Department of Geosciences, Penn State Astrobiology Research Center (PSARC), The Pennsylvania State University, University Park, PA, USA 2Department of Earth and Planetary Sciences, Harvard University, Cambridge, MA, USA 3Department of Biochemistry and Molecular Biology, The Pennsylvania State University, University Park, PA, USA 4Department of Earth Sciences, Indiana University-Purdue University Indianapolis, Indianapolis, IN, USA 5Department of Earth Sciences, University of California, Riverside, CA, USA 6Singapore Center for Environmental Sciences Engineering, Nanyang Technological University, Nanyang, Singapore 7Leibniz-Institut DSMZ-Deutsche Sammlung von Mikroorganismen und Zellkulturen, Braunschweig, Germany 8Department of Chemistry and Biochemistry, Montana State University, Bozeman, MT, USA

ABSTRACT

Mahoney Lake represents an extreme meromictic model system and is a valuable site for examining the organisms and processes that sustain photic zone (PZE). A single population of purple sulfur bacte- ria (PSB) living in a dense phototrophic plate in the chemocline is responsible for most of the primary pro- duction in Mahoney Lake. Here, we present metagenomic data from this phototrophic plate – including the genome of the major PSB, as obtained from both a highly enriched culture and from the metagenomic data – as well as evidence for multiple other taxa that contribute to the oxidative sulfur cycle and to sulfate reduction. The planktonic PSB is a member of the , here renamed Thiohalocapsa sp. strain ML1. It produces the okenone, yet its closest relatives are benthic PSB isolates, a finding that may complicate the use of okenone (okenane) as a biomarker for ancient PZE. Favorable thermodynamics for non-phototrophic sulfide oxidation and sulfate reduction reactions also occur in the plate, and a suite of organisms capable of oxidizing and reducing sulfur is apparent in the metagenome. Fluctuating supplies of both reduced carbon and reduced sulfur to the chemocline may partly account for the diversity of both autotrophic and heterotrophic . Collectively, the data demonstrate the physiological potential for maintaining complex sulfur and carbon cycles in an anoxic water column, driven by the input of exogenous organic matter. This is consistent with suggestions that high levels of oxygenic primary production maintain episodes of PZE in Earth’s history and that such communities should support a diversity of sulfur cycle reactions.

Received 24 January 2014; accepted 30 May 2014

Corresponding authors: T. Hamilton. Tel.: +18148635168; fax: +18148637823; e-mail: [email protected]; A. Pearson. Tel.: +16173848392; fax: +16178928340; e-mail: [email protected]; J. L. Macalady. Tel.: +18148656330; fax: +18148637823; e-mail: [email protected]

were common in the Mesoproterozoic (Canfield, 1998; INTRODUCTION Brocks et al., 2005; Reinhard et al., 2013). Although rare Euxinic conditions have been an intermittent feature of today, persistently euxinic environments serve as modern global oceans throughout Earth’s history. They accompa- analogues to ancient water columns and their associated nied significant events such as Phanerozoic biotic crises sedimentary basins. Of these modern analogues, meromic- (e.g., Pancost et al., 2004; Grice et al., 2005) and perhaps tic lakes have been of intense interest due to their

© 2014 John Wiley & Sons Ltd 451 452 T. L. HAMILTON et al. phototrophic chemocline communities, large geochemical sulfate-reducing supply sulfide for both photo- gradients, and laminated sedimentary records that can pre- trophic and chemolithoautotrophic primary productivity. serve (Meyer & Kump, 2008). Mahoney These primary producers in turn supply some of the Lake (ML) is a small meromictic lake in British Columbia, organic carbon necessary to support the heterotrophic sul- Canada, that represents an extreme end-member of such fate reducers, resulting in a coupled reductive and oxidative euxinia. The hypolimnion of the lake contains high levels sulfur cycle. Seasonal analyses of primary productivity and of sulfide (30–35 mM). ML has a pH in the range 7–9 and sulfur oxidation and reduction rates (Overmann et al., thus is not a soda lake (Northcote & Hall, 1983). Its dom- 1996a), and modeling of the carbon cycle in the inant anion is sulfate (400–500 mM), which is supplied by chemocline (Overmann, 1997), indicate that the system local surface waters draining alkaline lavas rich in Mg2+, must depend on additional organic carbon or other elec- 2+ + 2 2 Ca ,Na, and SO4 , and CO3 (Northcote & Hall, tron donors to sustain the cycle. Examining the microbial 1983). The hypolimnion is anoxic, eutrophic, and very taxonomic and functional potential of such systems is criti- saline (~74 g L 1 total dissolved solids (TDS)), while cal to understanding the links between aerobic and anaero- the epilimnion is oxic, oligotrophic, and of lower salinity bic biogeochemical feedbacks (Johnston et al., 2009). (~28 g L 1 TDS) (Northcote & Hall, 1983). Here, we evaluated the metabolic potential of the organ- Mahoney Lake (ML) is stable year-round due to density isms in this community using genomic and metagenomic stratification, and a 10- to 20-cm-thick chemocline occurs sequencing. High-throughput DNA sequencing provides at a depth of about 7 m. This layer supports a dense (~108 opportunities to obtain nearly complete genomes in systems cells mL 1) layer of anoxygenic phototrophs, limiting the of moderate complexity, or in systems with low strain diver- mixing of nutrients to the epilimnion and absorbing all vis- sity of a specific ecotype (e.g., Tyson et al., 2004; Iverson ible light (Overmann et al., 1991, 1996a, 1996b). Primary et al., 2012). In accordance with this, we were able to ana- productivity has largely been attributed to a single purple lyze several nearly complete genomes. We confirmed that sulfur bacterium (PSB), which was originally classified as a strain ML1 (Chromatiaceae) is the dominant PSB in the Ma- Thiocapsa sp. based on cell morphology (Northcote & Hal- honey Lake chemocline and that its closest taxonomic rela- sey, 1969). However, cultivation studies yielded an oke- tives belong to the genus Thiohalocapsa. In addition, there is none-producing member of the family Chromatiaceae, a single major taxon of Epsilonproteobacteria (Sulfurimonas denoted strain ML1. Okenone, a biomarker for some spe- sp.); eight putative sulfate-reducing Deltaproteobacteria; cies of PSB (e.g., Brocks & Schaeffer, 2008), is abundant other phototrophs including , Alphaproteo- in the lake sediments (Overmann et al., 1993; Coolen & bacteria, and GSB; and a diverse array of other organisms Overmann, 1998). Based on morphological and physiolog- from the phyla , Mollicutes, , Ac- ical properties, strain ML1 was reclassified as Amoebobacter tinobacteria, and Deinococcus. The results suggest that a purpureus (Overmann et al., 1991) and was subsequently high degree of functional redundancy exists within a system reclassified again to the genus Lamprocystis (Imhoff, 2001). that nonetheless possesses limited taxonomic diversity More recently, 16S rRNA and pufLM gene sequencing within the major populations. The data also point to specific revealed that strain ML1 is only distantly related to the interpretations about the energy balance of sulfide-driven type strain of Lamprocystis purpurea (Coolen & Overmann, autotrophy and – via the taxonomic placement of genes for 1998; Tank et al., 2009). okenone synthesis – address the continuity between plank- While small numbers of (GSB) are tonic and benthic low-oxygen ecosystems. also present in and below the chemocline, the total sulfide oxidation rate cannot be attributed to PSB and GSB alone MATERIALS AND METHODS (Overmann et al., 1996a). Chemoautotrophic sulfide oxi- dation has been observed near the oxic–anoxic interfaces of Chromatiaceae strain ML1: DNA Extraction, genomic other stratified systems and by analogy may be expected in sequencing, and annotation Mahoney Lake (Hadas et al., 2001; Buhring€ et al., 2011). Analysis of 16S rRNA gene sequences supports this, show- A sample for enrichment and isolation of strain ML1 was ing the presence of Epsilonproteobacteria that are likely to collected from the chemocline of Mahoney Lake at a depth be sulfide and/or sulfur oxidizers (Klepac-Ceraj et al., of 7-m on October 1, 1988 (Fig. 1A; J. Overmann, 2012). Sulfate reduction also occurs in the chemocline, unpubl. data), and has been maintained since as an and in situ generation is thought to supply much of the enrichment culture (Fig. 1B) using Pfennig’s medium as sulfide necessary to support anoxygenic described previously (Overmann & Pfennig, 1989). (Overmann et al., 1991, 1996a). Genomic DNA was extracted according to the cetyltrime- Collectively, these data indicate that complex cycling of thylammonium bromide (CTAB) protocol for bacterial carbon and sulfur occurs in the Mahoney Lake chemocline. DNA isolation of the Joint Genome Institute [http://my. The ML chemocline maintains a stable ecosystem in which jgi.doe.gov/general/]. Purified DNA was sequenced with

© 2014 John Wiley & Sons Ltd Sulfur cycle in a meromictic lake 453

A B pellet with an e.Z.N.A SP Maxi Kit (Omega Biotek, Norcross, GA, USA) according to the instructions of the manufacturer. The yield and quality of the extracted DNA was assessed using gel electrophoresis and spectrophotome- try (Nanodrop, Wilmington, DE, USA). Fragmentation and library preparation were performed by the North Carolina State University Genomic Sciences Laboratory. Paired-end 150-bp Illumina (HiSeq 2500; Illumina, San Diego, CA, µ 5 M 5 µM USA) sequencing was performed at the Harvard Center for Systems Biology and resulted in ~300 000 000 high-quality Fig. 1 Phase contrast photomicrograph of the Mahoney Lake chemocline reads. community (A) and of strain ML1 in culture with visible intracellular sulfur globules (B). Metagenome assembly, binning, and annotation a 454 pyrosequencer (GS FLX+; Roche, Branford, CT, Illumina sequencing reads were trimmed with Trimmomat- USA) in the Schuster Lab at The Pennsylvania State Uni- ic 0.20 (Lohse et al., 2012). Only sequences with at least versity. A total of 278 972 211 bases in 750 635 reads, 50 base pairs in both the forward and reverse direction with an average read length of 372 bp were generated and were retained and assembled into contigs with IDBA-UD assembled by Newbler assembler version 2.6 (Roche) into 1.09 (Table S1), using eight threads with default parame- 867 contiguous reads (contigs) of 500 bp or larger with ters (Peng et al., 2012). Coverage was determined by an average read depth of ~ 409. Contigs belonging to aligning raw reads to contigs with BWA 0.5.9 (Li & strain ML1 were assigned by their GC-content and verified Durbin, 2009). with BLAST (Altschul et al., 1990); 137 contigs The assembled contigs were binned using emergent (6 305 354 bp in total) with content of >53% were self-organizing maps (ESOM) based on tetranucleotide assigned to strain ML-1 in this way (Table 1). Contigs frequency, which resulted in clusters corresponding to were annotated using RAST (http://rast.nmpdr.org; Aziz taxonomically sorted tetranucleotide usage patterns (Dick et al., 2008). et al., 2009). A custom Python script (available at http://github.com/bovee/Ochre) was used to calculate tetranucleotide frequency of all contigs ≥2500 bp. Corre- Environmental sample collection, DNA extraction, and sponding reverse-complement tetranucleotides were com- metagenomic sequencing bined as described (Dick et al., 2009). Contigs were split Samples were collected from the phototrophic plate at 7 m into 5000-bp segments, clustered into taxonomic groups from Mahoney Lake, British Columbia (49°170N, [or ‘genomic bins’ (Voorhies et al., 2012)] by tetranucle- 119°350W) in July 2008. Samples were frozen on dry ice for otide frequency and visualized with Databionic-ESOM transport to Harvard University, where they were stored at (http://databionic-esom.sourceforge.net) – again using 80 °C until further processed. Samples were thawed and parameters from Dick et al. (2009) (Fig. S1). Bins were centrifuged, and DNA was extracted from the resulting inspected for coverage. (Multimodal distribution could indicate the bin contained scaffolds derived from multiple Table 1 Statistics for the strain ML1 draft genome and the ML-Gamma genomes, requiring that the bin be further split, but no and ML-Epsilon genomic bins such instances were found.) Following manual inspection and further curation by BLAST, -level taxonomic Strain ML1 ML-Gamma ML-Epsilon assignment was performed using Phyloshop (Shah et al., Scaffolds 137 485 72 2010) and MEGAN (Huson et al., 2011). Contigs larger Longest (bp) 332 978 105 972 114 097 than 2500 bp were annotated with an in-house General information Total bp 6 305 354 5 442 835 2 742 932 annotation pipeline as described in the supporting online Average length (bp) 46 024 12 597 29 155 material (SOM). N50 (bp) 114 369 17 120 38 096 The two highest-coverage bins in the metagenome, a Characteristics Gammaproteobacterium referred to here as ML-Gamma G + C content 68.3 65.9 38.6 and an Epsilonproteobacterium referred to as ML-Epsi- Protein coding genes 5664 5433 2732 rRNA genes lon, plus four Deltaproteobacteria bins (ML-Delta 1-4) 5S 1 1 0 (Figs S1 and S2), were selected for in-depth characteriza- 16S 1 1 0 tion and taxonomic assignment of their predicted genes. 23S 1 1 0 For the ML-Gamma and ML-Epsilon bins, paired reads tRNA genes 42 57 48 mapping to scaffolds from each bin were reassembled

© 2014 John Wiley & Sons Ltd 454 T. L. HAMILTON et al. using Velvet (Zerbino & Birney, 2008) as previously respectively; and the versions described in this paper are described (Hug et al., 2013). Scaffolds of each re-assem- AZMQ01000000 and AZMQ010000000, respectively. The bly were annotated as described in the SOM. Genome draft genome sequence of Thiohalocapsa sp. strain ML1 completeness was evaluated using a suite of 76 genes (Chromatiaceae) is deposited at DDBJ/EMBL/GenBank selected from a set of single-copy marker genes that show under the accession JABX00000000. The version described no evidence for lateral gene transfer (Sorek et al., 2007; in this paper is version JABX01000000. Gene IDs reported Wu & Eisen, 2008) (Tables S2 and S3). The in Table S6 represent the assembled gene sequences available ML-Gamma genomic bin was annotated by RAST for on the IMG/M Web site. comparison to the strain ML1 (Chromatiaceae) genome. The number of genomes per bin assigned to the RESULTS Deltaproteobacteria was estimated using 35 single-copy orthologous groups (Raes et al., 2007; Handley et al., Genomic sequencing of the highly enriched culture of 2012; Table S4). Genome coverage was estimated by strain ML1 resulted in a draft genome of strain ML1, a assuming that the genome size of each phylotype was Chromatiaceae sp. Contigs of the draft genome of strain approximately the same as its closest relative (Whitaker & ML1 were easily resolved from a contaminant organism Banfield, 2006; Jones et al., 2012). belonging to the class Mollicutes by %GC (68% vs. ~30%) and verified by BLAST. The draft genome assembled into 137 scaffolds containing 6.3 Mbp (Table 1). Taxonomic assignments and phylogenetic analyses De novo assembly of the metagenomic data from 7 m 16S rRNA sequences often fail to assemble into larger con- water depth resulted in 452 140 contigs containing ~ 640 tigs and scaffolds in metagenomic studies, and independent Mbp. The vast majority of contigs could be assigned to assemblies of these genes [e.g., Phyloshop (Shah et al., bacteria, while smaller numbers of eukaryotic, archaeal, and 2010) or EMIRGE (Miller et al., 2011)] are difficult to viral sequences were recovered (Fig. 2A). Most of the bac- assign to bins separated by genomic signatures such as terial contigs (>60%) were affiliated with , %GC-content or tetranucleotide frequency. To overcome but sequences affiliated with Firmicutes, Bacteroidetes, and these obstacles, single-copy ribosomal proteins were ana- were also abundant (Fig. 2B). Tetranucleo- lyzed as described in the SOM (Table S5) to make genus- tide frequency analysis of the de novo metagenomic assem- level assignments of the ML-Gamma genomic bin and the bly resolved 26 distinct taxonomic bins (Fig. S1), ML-Epsilon genomic bin. Such analyses yield resolution including well-defined genomic bins for a Gammaproteo- comparable to that for 16S rRNA phylogenetic trees (Hug bacterium (ML-Gamma) and an Epsilonproteobacterium et al., 2013). (ML-Epsilon). These two organisms represent the domi- The phylogenetic position of bacterial 16S rRNA genes nant members associated with the oxidative arm of the sul- was evaluated by approximate likelihood-ratio tests (Anisim- fur cycle. Four distinct bins of sulfate-reducing ova & Gascuel, 2006) as implemented in PhyML. Deltaproteobacteria (ML-Delta), including Desulfobacte- Sequences were aligned and the best evolutionary model rales and Desulfovibrionales, were also resolved, and these was determined using JMODELTEST (version 2.1.1, Darriba bins collectively contained at least eight genomes (Fig. S2). et al., 2012). Maximum-likelihood reconstructions used the As observed in other studies employing similar binning General Time Reversible substitution model and gamma- strategies (e.g., Handley et al., 2012), the coverage, GC- distributed rate variation with a proportion of invariable content, and phylogenetic similarity of scaffolds within the sites as recommended by JMODELTEST. Phylograms were Deltaproteobacteria bins precluded further resolution of rate-smoothed using the multidimensional version of Ram- individual genomes. In contrast, the Gamma- and Epsilon- baut’s parameterization as implemented in PAUP v. 4.0 proteobacteria bins each contained a single genome. We (Swofford, 2001) as previously described (Meuser et al., also recovered six low coverage bins affiliated with the 2013). Methods for gene-specific phylogenetic analyses Rhodobacteriales of the , three bins of (e.g., soxB, cruO; Table S6) are provided in the SOM. Actinobacteria and Bacteroidetes, two other (non-photo- trophic) bins (Xanthomonadales and Thiotricales), and single bins affiliated with Betaproteo- Nucleotide sequence accession numbers bacteria (Burkholderiales), Cyanobacteria, Spirochetes, and Metagenomic sequence data are deposited at CAMERA Mollicutes, (Fig. S1). Based on %GC and BLAST results, the (http://camera.calit2.net/), and the assembled sequences Mollicutes bin recovered from the metagenome is not the can be accessed via IMG/M (http://img.jgi.doe.gov). same Mollicutes organism present in the co-culture with Metagenome bin sequences – ML-Gamma and ML-Epsilon strain ML1 (data not shown). These results reveal a diverse – are deposited at DDBJ/EMBL/GenBank under the acces- capacity for sulfur and carbon cycling in the Mahoney Lake sion numbers AZMQ00000000 and AZMQ00000000, chemocline.

© 2014 John Wiley & Sons Ltd Sulfur cycle in a meromictic lake 455

99 Marichromatium purpuratum 984 (NR_037103) A 100 Marichromatium sp. L55 (EU878307) Marichromatium sp. JA349 (AM947937) bheemlicus JA132 (NR_042524) 98 Thiorhodococcus sp. AK35 (HF562219) Thiorhodococcus kakinadensis (NR_042526) 93 Thiorhodococcus drewsii AZ1 (FM178273) 100 vinosum DSM 180 (NR_074584) Allochromatium renukae JA136 (NR_042563) okenii DSM169 (NR_025315) 93 Thiocystis violascens DSM198 (NR_102951) 100 Lamprocystis purpurea DSM 4197 (NR_028885) Lamprocystis roseopersicina strain DSM 229 (NR_025288) Candidatus Thiodictyon syntrophicum str. Cad16 (AJ006061) Thiocapsa sp. BH-1 (AJ871143) 100 Thiocapsa roseopersicina strain 5811 (NR_041728) Thiocapsa marina strain BM-3 (AJ581994) mobilis 8321 (NR_102479) 100 Thiocapsa pfennigii 4250 (NR_036977) Thiococcus sp. AT2204 (AJ401211) 98 ML1 Thiohalocapsa halophila 4270 (NR_028863) glycolicum BN 3201 (NR_044896) 100 Halochromatium salexigens 6310 (NR_036810) Bacteria Halochromatium roseum JA134 (NR_042529) Thiohalocapsa marina JA142T (AM491592) Thiorhodovibrio winogradskyi strain DSM 6702T (Y12368) Eukarya 100 Thiorhodovibrio sp. 970 (FJ815159) 100 Ectothiorhodospira sp. PHS1 (EU590916) Virus 100 Ectothiorhodospira sp. AM4 (EU252492) Ectothiorhodospira variabilis (AM943122) 100 Thioalkalivibrio sp. K90mix (NR_074730) 94 Thioalkalivibrio sulfidophilus HL-EbGr7 (NR_074692) 94 Thioalkalivibrio denitrificans strain ALJD (NR_028745) B 0.01

Fig. 3 Maximum-likelihood-based phylogenetic 16S rRNA gene tree of closely related Gammaproteobacteria and strain ML1 (Chromatiaceae). Accession numbers are provided in parentheses. Gray circles indicate species that synthe- size okenone. Bootstrap support values based on 1000 bootstrap samplings >90 are noted.

taxa are sufficient to demonstrate a complete oxidative, reductive, and phototrophic sulfur cycle. We also discuss the correspondence between the high-coverage Gamma- Actinobacteria Alphaproteobacteria proteobacterium and isolate strain ML1, and we address Betaproteobacteria implications for the presence of the carotenoid pigment, Deltaproteobacteria Epsilonproteobacteria okenone, in the geologic record. Gammaproteobacteria Bacteroidetes Firmicutes Oxidative sulfur cycle Other Strain ML1 (Chromatiaceae) and the ML-Gamma genomic bin C ML-Gamma The highest-coverage organism in the metagenome, ML- ML-Epsilon Gamma (Fig. 2C), contains a 16S rRNA gene identical to strain ML1 (Coolen & Overmann, 1998; EMBL acc. no. AJ006212). Phylogenetic analysis of 19 concatenated phylo- genetic marker genes (Table S5) indicates that the ML- ML-Delta Gamma genome bin is strain ML1 or a very closely related ecotype (Fig. S3). Alignment of single-copy genes between strain ML1 and the ML-Gamma bin (Table S2) shows >99% identity across these phylogenetic markers. The ML-Gamma 0 200 400 600 800 Coverage genome bin is similar in size (5.7 Mbp) to the draft genome (6.3 Mbp) of strain ML1 and contains 73 of these 76 mar- Fig. 2 Taxonomic affiliation of all sequences in the 7-m metagenome and ker genes (Table S2), indicating that it is >90% complete. coverage of genome bins. (A) Number of sequences affiliated with each Taxonomic assignment places strain ML1 into the order . (B) Number of sequences assigned at the phylum level (except the Chromatiales and the family Chromatiaceae. In contrast to Proteobacteria which are represented by class). (C) Genomic bins distin- guished by read coverage. Other indicates all phyla represented by <800 the initial assignment based on morphological and physio- sequences. logical characteristics, our 16S rRNA gene sequence analy- sis indicates that this strain is most closely related to Here, we will focus on genomic insights as they relate to Thiohalocapsa spp. and Halochromatium spp. (Fig. 3). The sulfur and carbon cycling catalyzed by the high-coverage 16S rRNA gene of strain ML1 is 98% identical to that of Gammaproteobacterium, the Epsilonproteobacterium, and Thiohalocapsa halophila 4270T (DSM 6210T), which is its the members of the Deltaproteobacteria. Together, these closest cultured relative. In contrast, phylogenetic analysis

© 2014 John Wiley & Sons Ltd 456 T. L. HAMILTON et al. based on concatenated sequences of phylogenetic marker Scaffold 2 genes shows that strain ML1 is most closely affiliated with crtB crtI Scaffold 29 Thiorhodovibrio spp., including Thiorhodovibrio sp. 970 and the Mahoney Lake shoreline isolate Thiorhodovibrio wino- crtE crtF Scaffold 33 gradsky (Overmann et al., 1992) (Fig. S3); however, this crtU crtY crtC cruS cruO analysis is restricted to available complete genomic sequences, which currently includes no members of the CrtI Neurosporene genera Thiohalocapsa and Halochromatium. Strain ML1 and the ML-Gamma metagenomic bin both encode the genes necessary for photoautotrophic growth CrtY utilizing reduced sulfur compounds as the electron source for carbon dioxide reduction. These genomes contain a puf γ-Carotene operon encoding a type-2 photochemical reaction center CrtC comprised of the subunits PufL, PufM, and PufH, which O are identical in arrangement and sequence to the previously CruO 1'-Hydroxy-γ-carotene published puf operon of the isolate (Tuschak et al., 2005); O bacteriochlorophyll a biosynthesis genes; and genes (e.g., O CrtF csoS) for alpha-carboxysomes. Both genomes contain cbbLS 1'-Hydroxy-1',2'-dihydro-β,ψ-carot-4'-one O and cbbM genes, which encode Forms I and II of RuBis- O CO, the key enzyme in the Calvin–Benson–Bassham CrtU Thiothece-474 (reductive pentose phosphate) cycle. Genes encoding the O large and small subunit of RuBisCO form I, cbbL and cbbS, O Okenone are clustered with cso genes encoding carboxysome shell proteins; while the form II gene cbbM is located elsewhere Fig. 4 Predicted pathway for okenone synthesis of strain ML1 and ML- in the genome. The CbbL and CbbM sequences of strains Gamma based on genes identified in the genome and genome bin, respec- ML1 and the ML-Gamma genome are closely related to tively. Scaffold numbers refer to the location of predicted genes in the draft those of the species ‘Candidatus Thiodictyon syntrophi- genome of strain ML1. Proteins and the reactions they catalyze in the path- way for the biosynthesis of okenone, from Vogl & Bryant, 2011, are cum’ strain Cad16 from Lake Cadagno (Peduzzi , et al. summarized. 2012) (Fig. S4), which preferentially expresses Form I Ru- v 0 BisCO when CO2 assimilation rates are high (Storelli -ring of okenone, and the unique CruO 4/4 ketolase et al., 2013). Consistent with the observation of gas vesi- protein (Vogl & Bryant, 2011, 2012) (Fig. S6). Impor- cles both in the isolate and in samples collected from the tantly, the genes encoding CruO and the associated CruS chemocline (Overmann et al., 1991), we observed gvp enzyme in the ML-Gamma genome bin are the only copies genes which encode gas vesicle proteins. Both genomes of these diagnostic genes detected in the entire metage- also encode a cbb3-type heme-copper cytochrome oxidase, nome, establishing that this PSB taxon is probably the only which is capable of maintaining catalytic activity under source of okenone in the chemocline. micro-oxic conditions (Pitcher & Watmough, 2004), as The genomes of strain ML1 and ML-Gamma encode well as genes for a hox-type, bi-directional [NiFe]-hydroge- thiosulfate oxidation enzymes (Sox), a flavocytochrome c- nase and succinate dehydrogenase. Genes encoding a sulfide dehydrogenase (Fcc), and a D-type sulfide:quinone Type-1 NADH dehydrogenase (nuo), Mo-dependent oxidoreductase (Sqr) (Fig. 5, Fig. S7) (Gregersen et al., nitrogenase (nifHDK) (Fig. S5), and an Rnf-electron 2011). SoxA, B, X, Y, and Z comprise a multi-enzyme transport complex (Rhodobacter nitrogen fixation, rnf- complex of SoxXA, a heterodimeric c-type cytochrome; encoded) were also identified in both genomes. Rnf com- SoxB; and SoxYZ, a heterodimeric thiosulfate-binding plexes are widely distributed and are thought to mediate protein necessary for oxidation of thiosulfate to sulfate. electron transfer between NADH and ferredoxin (Biegel As in other , no soxCD genes were et al., 2011), and they may also play a role in nitrogen fix- found in either the isolate genome or the genome bin. In ation (Jeong & Jouanneau, 2000). other Chromatiaceae that lack soxCD – for example, The data also confirm that this taxon encodes all Allochromatium vinosum – sulfur/polysulfide globules enzymes necessary for biosynthesis of the carotenoid form in the periplasm as oxidation intermediates (Hensen okenone, an abundant compound in the Mahoney Lake et al., 2006; Frigaard & Dahl, 2009; Dahl et al., 2013). chemocline particulates, along the shoreline, and in sedi- In Alc. vinosum, reverse dissimilatory sulfite reductase ment cores (Overmann et al., 1991) (Fig. 4). These data (DsrAB) and other related Dsr proteins – including the include genes encoding an O-methyltransferase (CrtF), the transmembrane complex DsrMKJOP, the DsrEFH CrtY and CrtU proteins necessary for synthesis of the complex, and DsrC – are necessary for oxidation of the

© 2014 John Wiley & Sons Ltd Sulfur cycle in a meromictic lake 457

Sox- Dissimilatory Assimilatory Sulfurimonas gotlandica catalyzed oxidation of sulfate reactions reduced sulfur reduction Sulfurimonas denitrificans DSM 1251 Sulfate Sulfate Sulfurimonas sp. AST-10 Thiosulfate CysND SoxABXYZ Sat Sat Sulfurimonas autotrophica DSM 16294 APS APS Sulfate, ML-Epsilon AprAB polysulfides CysC SoeABC Thiovulum sp. ES Sulfite PAPS Thiosulfate Sulfuricurvum kujiense DSM 16994 SoxABCDXYZ DsrAB CysH Nitratiruptor sp. SB155-2 Sulfate 0 S , polysulfides Sulfite 0.04 CysJI Sqr Sir Sulfide Fig. 6 Maximum-likelihood phylogenetic tree of 14 concatenated single- Sulfide copy ribosomal proteins (Table S5) showing the taxonomic placement of the ML-Epsilon genomic bin. Bootstrap support values based on 1000 boot- Fig. 5 Predicted oxidative sulfur cycle in the Mahoney Lake chemocline strap samplings were >85 for each node. based on genes identified in the metagenome. Colors indicate the phyloge- netic affiliation of the predicted proteins. Red, Alphaproteobacteria; Yellow, Epsilonproteobacteria; Purple, Gammaproteobacteria; Green, green sulfur lithoautotrophic, sulfur-oxidizing aerobe that can use bacteria; Gray, Cyanobacteria; Blue, Betaproteobacteria; Orange, Bacteroi- elemental sulfur, sulfide, and thiosulfate as electron donors detes. Abbreviations: APS, Adenosine-50-phosphosulfate; PAPS, 3-Phospho- adenosine-50-phosphosulfate. (Inagaki et al., 2003). Based on the presence of phyloge- netic marker genes, the assembled genome is estimated to stored sulfur to sulfite (Dahl et al., 2008; Oliveira et al., be >90% complete, and it is similar in size (~2.7 Mbp) 2008; Grein et al., 2010). The DsrEFH complex, involved (Table 1) to that of several closely related species, includ- in sulfur transfer reactions, is encoded in genomes of sul- ing the genome of S. autotrophica (~2.1 Mbp). Similar to fur-oxidizing that also encode DsrAB but is characterized chemolithoautotrophic Epsilonproteobacte- absent in sulfate-reducing organisms (Sander et al., 2006; ria, the ML-Epsilon genome encodes the enzymes neces- Grimm et al., 2008). Genes encoding the reverse Dsr pro- sary for carbon fixation via the reverse TCA (rTCA) cycle, teins, including DsrEFH, are present in strain ML1 (and including ATP-dependent citrate lyase (acl), pyruvate: ML-Gamma), and their gene context is similar to the ferredoxin oxidoreductase (porAB), and 2-oxoglutarate: genomic context that is observed in Alc. vinosum (Grein ferredoxin oxidoreductase (oorAB) (Campbell et al., et al., 2010). 2006), but does not contain RuBisCO genes (cbbLS or Genes encoding homologues of key enzymes involved in cbbM). Unlike S. autotrophica, the ML-Epsilon genome dissimilatory oxidation of reduced sulfur and assimilatory encodes enzymes for the reduction of nitrate to N2, sulfate reduction are present (Fig. 5), including an adeno- including a periplasmic nitrate reductase (NapAB), a cyto- sine phosphosulfate reductase (APS reductase, ApsAB) and chrome cd1-dependent nitrite reductase (NirS), a nitric a sulfate adenylyltransferase (ATP-sulfurylase, Sat). In some oxide reductase (NorBC), and nitrous oxide reductase PSB, indirect oxidation of sulfite in the cytoplasm results (NosZ). These genes, together with the sox and sqr genes, from the activity of APS reductase and ATP-sulfurylase; suggest that nitrate-dependent sulfide oxidation can occur however, these enzymes are not required for sulfite oxida- under anoxic conditions (Simon & Klotz, 2013). Similar tion and are not universally found in sulfur-oxidizing prok- to other Sulfurimonas spp., the ML-Epsilon genomic bin aryotes (Dahl et al., 2013). The genomes also encode encodes a Mo-dependent nitrogenase and an ammonium homologues of the SoeABC sulfite-oxidizing enzyme com- transporter (Amt). It also contains cytochrome cbb3 oxi- plex, which is believed to oxidize sulfite presumably by dase and cytochrome bd quinol oxidase genes, both of transferring electrons to the electron transport chain via which are high-affinity terminal oxygen reductases capable the quinone pool (Dahl et al., 2013), as well as enzymes of functioning under micro-oxic conditions; formate dehy- involved in assimilatory sulfate reduction including adenylyl drogenase (fdhABC genes); and two [NiFe]-hydrogenases, sulfate kinase (CysC), phosphoadenosine phosphosulfate one cytoplasmic and one membrane-bound – suggesting reductase (CysH), and an NADPH-dependent sulfite this organism may use H2 and formate as electron donors reductase (CysJI) (Fig. 5). in addition to sulfide. Reflecting its role in the sulfur cycle, the genome The ML-Epsilon genomic bin contains a B-type sulfide:quinone oxidoreductase (Sqr) The ML 7-m metagenome contains a well-defined geno- (Fig. S7) (Gregersen et al., 2011), which catalyzes the mic bin identified as an Epsilonproteobacterium closely oxidation of sulfide to polysulfide, and soxCD genes, related to Sulfurimonas autotrophica (Fig. 6), a chemo- which indicate that the organism is capable of oxidizing

© 2014 John Wiley & Sons Ltd 458 T. L. HAMILTON et al. thiosulfate to sulfate (Fig. 5). The genome also encodes a organismal redundancy is also present within the reductive membrane-bound polysulfide reductase (psr-encoded), arm of the sulfur cycle. which may enable the use of polysulfide as an electron The genomic bins assigned to the Desulfurobacteriales acceptor (Hedderich et al., 1998); a sulfite cytochrome c and Desulfovibrionales encode the enzymes necessary for oxidoreductase (Sor), possibly involved in the inorganic dissimilatory sulfate reduction, including a sulfate adenylyl- disproportionation of reduced sulfur compounds (Finster, transferase (ATP-sulfurylase, Sat), an APS reductase 2008); and a rhodanese-related sulfurtransferase presum- (AprAB), and the dissimilatory sulfite reductase complex ably involved in the oxidation of thiosulfate (Ghosh & (DsrAB, DsrC, and DsrMKJOP), which catalyzes the reduc- 2 2 Dam, 2009). tion of SO3 to S (Fig. S8B). The genomic bins also encode Mn-dependent, inorganic pyrophosphatase (PpaC) Other organisms putatively involved in the oxidative and abundant ferredoxins (predominantly Ferrodoxin I). sulfur cycle The genomes contain homologues of D-lactate dehydroge- The 7-m metagenome contains multiple copies of key nase (gene ldhA) and pyruvate dehydrogenase (gene pdh) enzymes involved in sulfur oxidation that are assigned to for conversion of lactate to acetyl-CoA, as well as carbon other taxonomic bins. These include genes encoding monoxide dehydrogenase/acetyl-CoA synthetase (CODH/ sulfide:quinone oxidoreductase (Sqr), flavocytochrome c ACS). Many acetotrophic SRBs employ a CODH/ACS (Fcc), components of dissimilatory sulfite reductase pathway to cleave acetyl-CoA, and yet the presence of (DsrAB, DsrEFH, DsrC, and the DsrMKJOP complex), CODH/ACS genes in SRBs closely related to autotrophic the Sox multi-enzyme complex, and components of assimi- species (i.e., Desulfovibrio spp.; Fig. S8A) also suggests at latory sulfate reduction (CysND, CysC, CysH, CysJI, Sat, least facultative autotrophy may occur (Schauder et al., and Sir) (Fig. 5; Table S7). The sox-encoded enzymes of 1989). The ML-Desulfovibrionales genomic bins also encode thiosulfate oxidation are most abundant among these, fumarate reductases, indicating that these organisms may be while six copies each of homologs of DsrE, F, and H, spe- capable of fumarate respiration; a periplasmic [FeFe]-hydro- cific to sulfur-oxidizing organisms are present. Full-length genase (HydA); and heterodisulfide reductase-like proteins, SoxB sequences affiliated with the Alpha-, Beta-, and Gam- specifically HdrA- and HdrD-related proteins, which are maproteobacteria were recovered from other tetranucleo- thought to be involved in sulfite reduction through DsrC or tide-resolved genomic bins (Fig. 7). More putative reduction of the menaquinone pool (Pereira et al., 2011). sequences for SoxA and SoxB were present (31 and 33 The SRB genomic bins also encode Mo-dependent nitro- copies, respectively) than the sox-encoded proteins C, D, genases (nifHDK). X, Y, Z (Table S7), and the majority are affiliated with Al- phaproteobacteria. Several sox and sqr genes affiliated with Other organisms putatively involved in the reductive GSB, which use the Sox system for thiosulfate oxidation sulfur cycle (Ogawa et al., 2008; Gregersen et al., 2011), were also Genes from other species involved in reducing sulfur com- observed (Table S7), despite the low number of reads pounds include 21 unique copies of dsrD. Although the assigned to the phylum Chlorobi. The translated sqr exact function of DsrD remains unknown, it is postulated sequences cluster near the SqrD-type of Chlorobaculum to be involved in sulfite reduction (Grein et al., 2013) and tepidum TLS (Gregersen et al., 2011; Holkenbrink et al., is typical of SRB, but it is not found in sulfur-oxidizing 2011) (Fig. S7; Table S8). organisms (Grimm et al., 2008). The metagenome also contains tetrathionate reductase genes (ttr), which catalyze the reduction of tetrathionate to thiosulfate and are widely Reductive arm of the sulfur cycle distributed in the Enterobacteriaceae and other facultative ML Deltaproteobacteria community genomic bins anaerobes (Zopfi et al., 2004). Dimethyl sulfoxide The Mahoney Lake chemocline contains multiple taxa of (DMSO) reductase (DmsA), also common in anaerobic Deltaproteobacteria, including at least eleven distinct geno- heterotrophic bacteria, was identified. The majority of ttr types (Fig. S8A) based on ribosomal protein S3 sequences and dmsA sequences were affiliated with the Firmicutes, (Hug et al., 2013). At least six genomes of Desulfurobacte- Deltaproteobacteria, and the Alpha- and Gammaproteobac- riales spp. (present in genomic bins designated Delta 1, 3, teria (Table S7). and 4) and at least two Desulfovibrionales spp. genomes (genomic bin designated Delta 2) (Table S9, Figs S1 and Phylogenetic diversity of dsr S2) could be resolved. The high degree of phylogenetic similarity, GC-content, and coverage among the Deltapro- The 7-m metagenome contains eleven distinct copies of teobacteria spp. precludes the resolution of single ge- dsrAB, encoding dissimilatory sulfite reductase, the enzyme nomes. The recovery of multiple sulfate-reducing bacteria necessary for sulfite reduction or sulfur oxidation (Fig. 8). (SRB) genomes across several tetranucleotide bins indicates Concatenated amino acid sequence alignments of full-

© 2014 John Wiley & Sons Ltd Sulfur cycle in a meromictic lake 459

100 Rubellimicrobium thermophilum DSM 16684 7m SoxB_1 7m SoxB_2 Roseovarius sp. 217 Thalassobacter arenae DSM 19593 100 Roseobacter sp. AzwK-3b 7m SoxB_3 100 Maritimibacter alkaliphilus 100 Sulfitobacter sp. NAS-14.1 ATCC 17025 100 100 7m SoxB_4 Dinoroseobacter shibae DFL 12 Xanthobacter autotrophicus Py2 Magnetococcus marinus MC-1 100 Sulfuricella denitrificans Sideroxydans lithotrophicus ES-1 Thiobacillusthioparus 100 Thiobacillus denitrificans ATCC 25259 Thauera terpenica 58Eu Hydrogenophaga sp. PBC Cupriavidus sp. BIS7 Ralstonia eutropha H16 100 Herminiimonas arsenicoxydans 7m SoxB_5 Thiomonas sp. 3As Polaromonas sp. JS666 100 Polaromonas sp. CF318 100 Bradyrhizobium sp. STM 3809 100 Nitrobacter hamburgensis X14 95 Afipia sp. 1NLS2 Methylobactermarinus 100 Thiomicrospira halophila 91 7m SoxB_6 100 Thiomicrospira arctica 7m SoxB_7

100 Thioalkalimicrobium cyclicum ALM1 93 Thioalkalimicrobium aerophilum Thiocystis violascens DSM 198 Lamprocystis purpurea DSM 4197 Thiocapsa marina 5811 95 ML1 SoxB 100 Thiorhodococcus drewsii Marichromatiumpurpuratum 987 Allochromatium vinosum DSM 180 Neptuniibacter caesariensis 98 Beggiatoa sp. PS Beggiatoa alba 100 Chlorobium chlorochromatii CaD3 Pelodictyon phaeoclathratiforme BU-1 100 Chlorobium limicola TLS Fig. 7 Phylogenetic diversity of SoxB in the 100 Chlorobium tepidum Mahoney Lake chemocline. Maximum- 85 Chlorobium phaeovibrioides DSM 265 likelihood phylogenetic tree constructed from 100 Chlorobaculum parvum NCIB 8327 translated SoxB sequences identified in the ML Acidithiobacillus thiooxidans 7-m metagenome. Red, Alphaproteobacteria; 100 Sulfurimonas autotrophica DSM 16294 100 orange, Betaproteobacteria; purple, Gammapro ML-Epsilon SoxB 100 sp. GD1 teobacteria; green, green sulfur bacteria; black, 96 Sulfurimonas Thiomicrospira denitrificans Acidithiobacillia; yellow, Epsilonproteobacteria; 100 Sulfuricurvum kujiense DSM 16994 blue, Aquificales. The strain ML1 genome and Sulfurovum sp. NBC37-1 ML-Gamma genomic bin contain identical SoxB Persephonella marina EX-H1 sequences represented here as ML1. 100 Sulfurihydrogenibium sp. YO3AOP1 Abbreviations: SRB, sulfate-reducing bacteria; 100 100 Hydrogenobacter thermophilus TK-6 SOB, sulfur-oxidizing bacteria. Bootstrap Thermocrinis albus DSM 14484 support values based on 1000 bootstrap 100 Aquifex aeolicus VF5 samplings ≥85 are shown for each node. 0.2

© 2014 John Wiley & Sons Ltd 460 T. L. HAMILTON et al.

Thioalkalivibrio sp. ALJ17 Thioflavicoccus mobilis 8321 Lamprocystis purpurea DSM 4197 Thiorhodococcus drewsii AZ1 Allochromatium vinosum DSM 180 Thiocystis violascens DSM 198 Thiorhodovibrio sp. 970 ML1 DsrAB Thiocapsa marina 5811 Marichromatium purpuratum 984 Thioalkalivibrio nitratireducens DSM 14787 7m DsrAB_1 7m DsrAB_2 Rhodomicrobium vannielii ATCC 17100 SOB halophila SL1 Magnetospirillum magneticum AMB-1 Magnetospirillum magnetotacticum Magnetospirillum sp. SO-1 Magnetococcus marinus MC-1 Chlorobium phaeobacteroides BS1 Chlorobium chlorochromatii CaD3 Chlorobium phaeovibrioides DSM 265 Chlorobium luteolum DSM 273 Pelodictyon phaeoclathratiforme BU-1 Chlorobaculum parvum NCIB 8327 Chlorobium tepidum TLS Prosthecochloris aestuarii DSM 271 Chlorobium limicola DSM 245 Desulfobulbus propionicus DSM 2032 Desulfobulbus rhabdoformis Desulfurivibrio alkaliphilus AHT2 Desulfocapsa sulfexigens DSM 10523 Desulfotalea psychrophila LSv54 7m DsrAB_3 7m DsrAB_4 7m DsrAB_5 Thermodesulforhabdus norvegica Thermodesulfatator indicus DSM 15286 Thermodesulfobacterium geofontis OPF15 Candidatus Desulforudis audaxviator MP104C delta proteobacterium NaphS2 Desulfotomaculum kuznetsovii DSM 6115 Desulfotomaculum thermocisternum Desulfotomaculum acetoxidans DSM 771 Desulfotomaculum gibsoniae DSM 7213 Desulfomonile tiedjei DSM 6799 Desulfatibacillum alkenivorans AK-01 7m DsrAB_6 Fig. 8 Phylogenetic diversity of dissimilatory Desulfococcus oleovorans Hxd3 SRB sulfate reductase (Dsr) in the Mahoney Lake Desulfobacterium autotrophicum HRM2 Desulfobacter postgatei chemocline. Maximum-likelihood phylogeny Desulfobacter curvatus calculated from a concatenated amino acid Desulfobacula toluolica Tol 2 7m DsrAB_7 sequence alignment of DsrA and DsrB 7m DsrAB_8 Desulfotignum phosphitoxidans sequences recovered from the ML 7-m 7m DsrAB_9 metagenome and representative sequences 7m DsrAB_10 Desulfonatronospira thiodismutans from the Deltaproteobacteria and sulfur- Desulfohalobium retbaense DSM 5692 oxidizing members of the Alpha- and Desulfomicrobium baculatum DSM 4028 Desulfovibrio hydrothermalis AM13 Gammaproteobacteria. DsrA and DsrB Desulfovibrio salexigens DSM 2638 Bilophila wadsworthia sequences of strain ML1 and ML-Gamma are Desulfovibrio alaskensis G20 identical and represented by ML1. Desulfovibrio desulfuricans Desulfovibrio africanus str. Walvis Bay Abbreviations: SRB, sulfate-reducing bacteria; Desulfovibrio magneticus RS-1 0.06 SOB, sulfur-oxidizing bacteria. length alpha and beta subunits (DsrA and DsrB, respec- encoding ribulose 1,5-bisphosphate carboxylase/oxygenase tively) confirm the presence of multiple sulfate-reducing (RuBisCO). Of these, five group with the Type IV RuBis- Deltaproteobacteria-affiliated sequences and several sulfur- CO-like proteins (RLP), which apparently do not catalyze oxidizing Alpha- and Gammaproteobacteria-affiliated RuBP carboxylation or oxygenation (Hanson & Tabita, sequences. Several of these SRB-affiliated sequences are not 2003; Tabita et al., 2007) (Fig. 9, Fig. S9). In the GSB represented in the genome bins discussed above and pre- Chlorobaculum tepidum, a RLP has been implicated in sulfur sumably are derived from other less abundant taxa in the (Hanson & Tabita, 2003), while in some Alpha- system. No dsrA or dsrB genes affiliated with GSB were proteobacteria, RLP are involved in 5-methylthioadenosine- recovered despite the observation that nearly all GSB con- dependent growth (Tabita et al., 2007). The remaining tain dsr (Frigaard & Bryant, 2008). This is most likely due sequences all branch with characterized RuBisCO proteins, to the low abundance of reads assigned to GSB in the 7-m indicating they are involved in CO2 assimilation in situ.In metagenome. addition to the Type I and Type II RuBisCO identified in the ML-Gamma genomic bin, the metagenome contains three sequences affiliated with Type I RuBisCO (similar to Phylogenetic diversity of RuBisCO Thioalkalvibrio spp. and members of the phyla Cyanobacte- In addition to the ML-Gamma cbbL and cbbM sequences, ria and Chloroflexi) and two affiliated with Type II RuBis- the 7-m metagenome contains fourteen full-length genes CO (similar to sequences of Sulfuricella denitrificans,a

© 2014 John Wiley & Sons Ltd Sulfur cycle in a meromictic lake 461

5 6 Type III 7 Hybrid Type II/III 4 8

3

9 95 97 ML1 CbbM ML1 CbbL2 100 100 92

Fig. 9 Phylogenetic diversity of RuBisCO in Type II the Mahoney Lake chemocline. Maximum- 1 likelihood phylogenetic trees constructed from Type I CbbL and CbbM sequences mined from the Type IV ML 7-m metagenome and the NCBI and RuBisCO-like IMG/M database. Numbers indicate full- length sequences mined from the ML 7-m metagenome (Table S6) and the full tree is shown in Fig. S9. CbbL and CbbM sequences of strain ML1 and ML-Gamma are identical 10 and represented by ML1. Bootstrap support 13 0.08 values based on 1000 bootstrap samplings are 14 12 11 shown for each node. sulfur-oxidizing Betaproteobacterium, and R. rubrum) vation bolstered by more recent analyses of the 16S rRNA (Fig. S9). Two additional sequences branch with the metha- and the pufLM genes, both of which indicated that strain nogen Type III RuBisCO and two sequences group ML1 is affiliated with the genus Thiohalocapsa (Tank et al., together with the novel bacterial–archaeal hybrid TypeII/ 2009). Phylogenetic analyses presented here and elsewhere III sequences (Fig 8), which are putatively functionally anal- (Coolen & Overmann, 1998; Tank et al., 2009) show this ogous to Type III (Tabita et al., 2007); these enzymes are organism is most closely related to marine mat-forming thought to fix CO2 and contribute to adenosine monophos- members of the Chromatiaceae, such as Thc. halophila as phate (AMP) recycling (Wrighton et al., 2012). well as Halochromatium roseum and Hch. glycolicum. The 16S rRNA sequences from Thc. halophila (formerly Thiocapsa halophila; Caumette et al., 1991) and strain DISCUSSION ML1 are 98% identical. Although some characteristics of strain ML1 – for exam- Identification of the dominant PSB in Mahoney Lake ple, the presence of gas vesicles and production of okenone The enrichment culture strain ML1 (Chromatiaceae; Over- – are also characteristic of Lamprocystis spp., strain ML1 mann et al., 1991) is very closely related to the highest- exhibits higher salt tolerance than freshwater species coverage population in the 7-m metagenome, indicating (Imhoff, 2005). Moreover, gas vesicles and okenone are not that strain ML1 is representative of the major PSB popula- taxonomically diagnostic. Gas vesicles have been observed in tion. This PSB is believed to be the most abundant organ- multiple genera of Chromatiaceae, including Thiocapsa, ism in the Mahoney Lake chemocline (Overmann et al., Thiodictyon, Thiolamprovum, Thiopedia, Lamprocystis, 1991), and the similarity between our metagenome (sam- , and Halochromatium (Kumar et al., 2007). pled in 2008) and strain ML1 (enriched in 1989) confirms Okenone production is similarly irregularly distributed; it is decadal stability of this population. The initial classification mainly found in the gas vesicle containing genera Thiocapsa, of strain ML1 as Amoebobacter purpureus was based on cell Thiopedia, Thiodictyon, Lamprobacter, and Thiolamprovum, morphology, the presence of gas vesicles, okenone produc- but it is also produced by the halophilic species Hch. roseum tion, and its ability to use glucose, pyruvate, and acetate (Kumar et al., 2007) and the non-gas vacuolate relative of (Overmann et al., 1991). Due to the re-classification of strain ML 1, Thc. halophila. Based on the robust 16S rRNA A. purpureus as a Lamprocystis species (Guyoneaud et al., gene analyses presented here, the clade containing strain 1998), strain ML1 has been referred to as ‘Lamprocystis ML1 and Thc. halophila is distinct from other major clades purpurea ML1’ (EMBL acc. nos. AJ006212, AY177752, of planktonic PSB (e.g., Lamprocystis spp., Chromatium Klepac-Ceraj et al., 2012). However, initial analyses of the spp.) (Fig. 3), including ‘Ca. Thiodictyon syntrophicum’ 16S SSU rRNA sequence suggested that the isolate had Cad16, an abundant PSB that occurs in the compact chemo- been misclassified (Coolen & Overmann, 1998), an obser- cline of Lake Cadagno (Storelli et al., 2013). As a result, we

© 2014 John Wiley & Sons Ltd 462 T. L. HAMILTON et al. suggest that strain ML1 is a potentially new species within settled and was captured by the dense phototrophic plate. the genus Thiohalocapsa and propose that it be renamed Both Cyanobacteria and Epsilonproteobacteria are capable Thiohalocapsa sp. strain ML1. of living in micro-oxic as well as sulfidic conditions (Jor- gensen et al., 1986; Campbell et al., 2006). Conversely, the bottom of the plate intersects the highly sulfidic A coupled sulfur cycle in a stable planktonic community (35 mM) hypolimnion, an environment also rich in polysul- Metagenomic sequencing of the Mahoney Lake chemo- fides, elemental sulfur, and other sulfur intermediates that cline indicates this sulfur-rich ecosystem harbors functional may be amenable to sulfur disproportionation. Single gen- redundancy in both the oxidative and reductive arms of ome bins (single populations) of sulfur oxidizers recovered the sulfur cycles (Fig. 10A). This functional redundancy in the present study contain multiple pathways for oxida- could be a key factor in maintaining the apparent stability tion of sulfur compounds, implying flexibility and/or of the Mahoney Lake chemocline over the past ~30 years, redundancy in their sulfur metabolism, a genomic charac- or it may have developed as a result of such stability. The teristic observed in other sulfidic systems (Jones et al., recovery of aerobic taxa such as Cyanobacteria and Epsi- 2012). Isotope values of sulfur species in the chemoclines lonproteobacteria may suggest fine-scale niche separation of other meromictic lakes, such as Fayetteville Green Lake, within the narrow phototrophic plate, the top of which is indicate that multiple processes contribute to sulfur bounded by the oxic epilimnion. Alternatively, these taxa oxidation and that these vary seasonally (Zerkle et al., may represent biomass from the oxygenated layer, which 2010). A similar seasonal variation may also occur in the

ABStrain ML1/ML-Gamma

cbb3 Dsr SO 2– CbbLS 4

CbbM Dsr S

alpha- o

LH2 carboxysomes Apr x SOB biomass TOC Calvin LH1 Sat 2– cycle Soe S O RC BChl a 2 3

Okenone f

c c Nif Hyd Gvp 0 SOB Rnf S N 2 ML-Gamma >>> NO – ML-Epsilon 3 NH ML-Epsilon, GSB, 3 NapAB Alpha-, Betaproteobacteria NO – Amt Fdh 2 cbb SO 2– Nir AclB 3 7 m 4 NO NorBC N O Hyd rTCA H S 2 2 NosZ Pili Nif Rhd SO 2– N SOR 4 SRB 2 Hyd ML-Delta spp.

Psr Sox SQR Chemocline (10 – 20 cm) H S, S O 2– 2 2 3 ML-Delta N 2 Sdh Hyd 2– Pdh Ldh H S SO

Nif 2 4

c t Dsr TCA y H S 3 2 0 Apr S Dsr Sat Nqr Hme Rnf Qmo Qrc SO 2– 4

Fig. 10 Genomic models of the strain ML1/ML-Gamma, ML-Epsilon and the ML-Delta spp. (A) and putative model of the sulfur cycle in the Mahoney Lake chemocline (B). The loss of biomass (organic carbon and reduced sulfur, purple arrow) from the chemocline due to rafting, external organic carbon as a sig- nificant source of carbon to SRB in the chemocline (TOC, brown arrow) and loss of S0 to the lower layers of the lake (yellow arrow with dashes) are depicted. Further results and discussion of strain ML1/ML-Gamma, ML-Epsilon, and the ML-Delta spp. are provided in the SOM. AclB, ATP-dependent citrate lyase;

Amt, ammonia/ammonium transporter; Apr, adenylyl sulfate (APS) reductase; cbb3, cytochrome c oxidase; Cyt3, cytochrome c3; Dsr, dissimilatory sulfite reductase; Fdh, formate dehydrogenase; Gvp, gas vesicle proteins; Hyd, hydrogenase; Hme, heterodisulfide reductase-like menaquinol-oxidizing enzyme com- plex; LH, light-harvesting complex; Ldh, lactate dehydrogenase; NapAB, NO3 reductase; Nif, nitrogen fixation; NirS, NO2 reductase; NorBC, NO reductase; + NosZ, N2O reductase; Nqr, Na -translocating NADH:quinone oxidoreductase; RC, photosynthetic reaction center; Pdh, pyruvate dehydrogenase; Psr, polysul- fide reductase; rTCA, reductive TCA cycle; Qmo, quinone-interacting membrane-bound oxidoreductase complex; Qrc, quinone reductase complex; Rhd, rho- danese-related sulfurtransferase; Rnf, Rnf-electron transport complex; Sat, ATP-sulfurylase; Sox, sulfur oxidation proteins; Sdh, succinate dehydrogenase/ fumarate reductase; Soe, sulfite-oxidizing enzyme complex; SOR, sulfite oxidoreductase; Sqr, sulfide:quinone oxidoreductase; SOB, sulfur-oxidizing bacteria; SRB, sulfate-reducing bacteria.

© 2014 John Wiley & Sons Ltd Sulfur cycle in a meromictic lake 463 chemocline of Mahoney Lake, supporting functional 1998), are also observed both diurnally and seasonally in redundancy in sulfur oxidation pathways and a diversity of Mahoney Lake. The associated small vertical excursions of sulfur-oxidizing species. the chemocline (Hall & Northcote, 1990) may help main- Historical data from Mahoney Lake show that the tain a diverse population of SRBs. Alternatively, varying chemocline is relatively stable in both thickness and depth, levels of oxygen or sulfide tolerance among taxa could although seasonal variation (e.g., autumnal die-off) affects result in increased diversity (Baumgartner et al., 2006). the total cell abundance of Thiohalocapsa sp. strain ML1 Encompassing both of these ideas, the metagenomic (Overmann et al., 1994). Other anaerobic phototrophic data presented here contain genetic evidence for both bacteria, including Thiocapsa sp. and GSB (Overmann heterotrophic and autotrophic SRB, implying that fluctua- et al., 1991) as well as other autotrophs (e.g., ML-Epsilon; tions in carbon supply play a role in maintaining multiple this study) and a large number of Alphaproteobacterial SRB spp. in the chemocline. SRB utilize organic sequences (this study), are observed in the Mahoney Lake compounds as electron sources for cellular metabolism. chemocline. However, in past studies, Thiohalocapsa sp. However, autotrophic growth of hydrogen-utilizing SRB strain ML1 was responsible for >98% of primary productiv- via the Wood-Ljungdahl pathway has been demonstrated ity and, on average, 75% of the sulfide oxidation (Over- in some species, including close relatives of some of those mann et al., 1991, 1996a). In contrast, the narrow recovered in the present study, that is, Desulfobacteriaceae chemocline of Lake Cadagno, a meromictic lake in Swit- (Rabus et al., 2013). The strong density gradient in Ma- zerland, hosts multiple populations of PSB, including honey Lake inhibits much of the in situ organic matter of Chromatium, Lamprocystis, Thiocystis, and Thiodictyon spp., the chemocline from entering the layers below (Overmann as well as GSB of the genus Chlorobium (Bosshard et al., et al., 1991), implying it is consumed by 2000). These multiple species of PSB persist and remain within the chemocline or lost to vertical mixing and/or active despite a recent increase in the abundance of Chloro- buoyancy. During the autumnal die-off and upwelling of bium clathratiforme (Decristophiris et al., 2009; Gregersen PSB biomass (Overmann et al., 1994), however, there is et al., 2009). In particular, ‘Ca. Thiodictyon syntrophi- minimal concomitant change in net activity in the cum’ strain Cad16 accounts for over 25% of primary pro- chemocline. This implies that the total chemocline sulfate duction and is also a prominent source of CO2 assimilation reduction rate is relatively stable, even if the loss of this in the dark (Storelli et al., 2013). organic matter affects the fractional balance of processes It is unclear what factors support the limited diversity (autotrophic vs. heterotrophic carbon supply). This among photoautotrophs in the Mahoney Lake chemocline. seasonal variation in organic carbon supply may support Several lines of evidence indicate that of the total Thiohalo- the multiple SRB spp. observed in the Mahoney Lake capsa sp. strain ML1 biomass, a relatively small proportion chemocline. (~15%), is physiologically active (Overmann et al., 1991, 1994). The formation of buoyant aggregates of Thiohalo- The energetic maintenance of euxinia capsa sp. causes high cell numbers to accumulate within the chemocline (Overmann & Pfennig, 1992), and the Although sulfate reduction and sulfide oxidation both high salinity of the hypolimnion and gas vacuolation of the occur in the Mahoney Lake chemocline, such a system is cells inhibit sedimentation of these aggregates (Overmann not self-sustaining. To satisfy the basic principles of et al., 1991). The elevated levels of sulfate and especially thermodynamics and maintain a coupled oxidative and sulfide (35 mM) in Mahoney Lake may create a very nar- reductive sulfur cycle, exogenous input of electrons is a row niche space that precludes the majority of anoxygenic stoichiometric requirement (Overmann, 1997). All of the photoautotrophs. Specific adaptations including elevated sulfide in Mahoney Lake must be derived from sulfate salt tolerance and adaptation to low light (Overmann reduction, with the electrons provided by respiration of et al., 1994) may enable Thiohalocapsa sp. strain ML1 to organic matter. However, oxygenic primary production in persist, albeit with very slow growth rates (Overmann the epilimnion is not sufficient to supply these electrons in et al., 1991, 1994). the form of phytoplanktonic carbon fixed in situ. Instead, Phylogenetic analyses of the ribosomal protein S3 indi- most of the needed organic matter must come from the cates multiple SRB spp. are present (Fig. S8A) in the Ma- surrounding landscape (Overmann et al., 1996a; honey Lake chemocline. This record of SRB is consistent Overmann, 1997). with the observation that sulfate reduction provides the Flux balance analysis confirms this principle. From data majority of reducing equivalents for PSB-dominated pri- in Overmann et al. (1996a), we calculate an average diffu- 2 1 mary production, and more specifically, that some of the sive vertical mixing of 9.2 mmol H2Sm day and 2 2 1 sulfide is supplied in situ rather than by vertical mixing 23 mmol SO4 m day into the chemocline (Overmann et al., 1996a). Redox oscillations at the oxic- (described in the SOM). The H2S flux is only sufficient to anoxic boundary, as is common in sediment systems (Aller, supply ~40% of measured oxidation rates (Overmann et al.,

© 2014 John Wiley & Sons Ltd 464 T. L. HAMILTON et al.

2 1991), implying that SO4 must be reduced in situ in the restricted – requiring micro-oxic or anoxic conditions, free chemocline to supply the remainder of the H2S. This is sulfide, and sunlight – diagenetic products of the caroten- consistent with suggestions that input of allochthonous – oid okenone have been invoked as markers for shallow and therefore necessarily oxygenic – organic matter is photic zone euxinia (Brocks et al., 2005; Brocks & Schaef- required to supply the extra flux of electrons needed to fer, 2008). However, the synthesis of okenone is irregularly sustain sulfate reduction (Overmann, 1997). The question distributed among Chromatiaceae, both taxonomically and then remains: where does all the upwelled sulfur go? The physiologically. Therefore, it is still debated whether diage- 2 epilimnion maintains a net SO4 deficit (and zero H2S) netic okenane in geologic sediments is of planktonic origin, relative to the hypolimnion, while sediment trap data show or if it also may be derived from benthic mat communities that <0.5% of the total upward sulfur flux sinks out of the (Meyer et al., 2011) in which PSB are widespread and chemocline as S0 (Overmann et al., 1994). To close mass diverse (Imhoff, 2005). Mat- or benthic-derived isolates balance, the remainder must be rafted to the shoreline in that produce okenone include Halochromatium roseum association with the annual partial die-off and buoyant dis- JA134 (Kumar et al., 2007), Thiocapsa marina (Caumette persal of the PSB from the chemocline (Overmann et al., et al., 1985, 2004), and the closest relative of Thiohalo- 1994)(Fig. 10B). This relationship implies that the net sul- capsa sp. strain ML1, Thc. halophila (Caumette et al., 2 2 0 fur cycle in the lake primarily is SO4 ? S (H2S) ? S , 1991; Kumar et al., 2007). that is, the terminal sink for sulfur is significantly more Okenone in ML sediments (Overmann et al., 1993; Co- 2 reduced relative to its oxidized input (SO4 from sur- olen & Overmann, 1998) has been linked to the presence rounding rocks). The ability to sustain net sulfate reduc- of Thiohalocapsa sp. strain ML1 in the water column tion in such a euxinic system, both in the chemocline and (Overmann et al., 1991). The metagenomic sequencing in the deeper hypolimnion, at first appears remarkable. and phylogenetic analyses presented here show that the However, simple free energy calculations under in situ only detectable okenone producer in the water column is conditions (DG = DG0 + RT lnQ) based on collated car- indeed this species. It is not only closely related to Thc. bon and sulfur species concentrations from the literature halophila, but it also appears to be more closely related to (Northcote & Hall, 1983; Overmann et al., 1991, 1996a, the Trv. winogradsky isolated from the shoreline mats b) show that while Mahoney Lake represents an extreme (Overmann et al., 1992) than to typical planktonic mem- end-member of euxinia with regard to sulfide concentra- bers of the Chromatiaceae, such as ‘Ca. Thiodictyon 2 tions, the H2S:SO4 ratio in this system still is not high syntrophicum’. The taxonomic similarity between Thiohalo- enough to thermodynamically inhibit sulfate reduction capsa, Halochromatium, and Thiorhodovibrio genera (Tank (Amend & Shock, 2001; Table S10, Fig. S10). The persis- et al., 2009; this work) suggests that a more general conti- tence of sulfate reduction under the extreme euxinic condi- nuity between planktonic and benthic PSB communities tions Mahoney Lake implies that – at least under exists, especially where oxygen minimum zones impinge on oxygenated atmospheres – SRB should be active in anoxic sediments within the photic zone. Similar to Mahoney systems throughout Earth history, regardless of the degree Lake, the okenone-rich chemocline of meromictic Fayette- of euxinia. ville Green Lake in New York hosts a diverse population of The redox stability and stable, poised conditions of PSB, and phylogenetic analyses of clones from benthic and Mahoney Lake likely result from its unique combination of planktonic habitats in that lake suggest these niches host high salinity, high sulfur concentrations, and high allochth- distinct PSB populations (Meyer et al., 2011). Okenone- onous carbon loading. The more dilute conditions of Lake rich sediments derived from benthic mats underlying the Cadagno – another analogue for ocean euxinia – result in oxic water column at Fayetteville Green Lake have chal- greater seasonal changes in fluxes, redox variability, and lenged the current interpretation of okenone as an exclusive microbial community succession (Bosshard et al., 2000; biomarker for planktonic PSB. In Mahoney Lake, the exam- Del Don et al., 2001; Decristophiris et al., 2009; Greger- ple of okenone production in the water column by a PSB sen et al., 2009). More generally, the concentration of free strain that is a close relative of benthic species further high- sulfide in any water column should scale directly with the lights the difficulty in linking okenone uniquely to water extent of organic matter overloading, and the extreme case column processes. While it remains assured that the recov- of Mahoney Lake undoubtedly is a factor in its biogeo- ery of okenane from ancient sediments is strong evidence chemical and ecological stability. for sunlit, sulfidic environments inhabited by PSB, the pres- ence of okenane alone may not be sufficient evidence for photic zone euxinia in the open water column. Implications for biosignatures of photic zone euxinia At the same time, in environments such as Mahoney The chemocline of Mahoney Lake provides valuable insight Lake where planktonic PZE clearly is present, chemoauto- into the geologic record of anoxygenic phototrophy. trophic sulfide oxidation and sulfate reduction also should Because the habitat of phototrophic sulfur bacteria is be suspected as co-occurring reactions in an active sulfur

© 2014 John Wiley & Sons Ltd Sulfur cycle in a meromictic lake 465 cycle, mediated by a wide variety of micro-organisms. complex, a novel ion-motive electron transport complex in This is distinct from a cryptic sulfur cycle in which no prokaryotes. Cellular and Molecular Life Sciences 68, 613–634. € net sulfur cycle redox processes occur (Canfield et al., Bosshard PP, Santini Y, Gruter D, Stettler R, Bachofen R (2000) Bacterial diversity and community composition in the 2010), but rather is an expected consequence of the chemocline of the meromictic alpine Lake Cadagno as water column anoxia associated with high organic carbon revealed by 16S rDNA analysis. FEMS Microbiology 31, fluxes. Under this model, both PZE and water column 173–182. sulfate cycling necessarily depend on oxygenic photosyn- Brocks JJ, Love GD, Summons RE, Knoll AH, Logan GA, thesis to supply electrons. This implies that PZE is only Bowden SA (2005) Biomarker evidence for green and purple sulphur bacteria in a stratified Palaeoproterozoic sea. Nature possible in high productivity, low and moderate pO2 437, 866–870. oceans (including ocean basins with restricted circulation). Brocks JJ, Schaeffer P (2008) Okenane, a biomarker for purple Episodes of euxinia in Earth history should require a con- sulfur bacteria (Chromatiaceae), and other new carotenoid fluence of all of these conditions, perhaps helping to derivatives from the 1640 Ma Barney Creek Formation. 72 – explain their apparent rarity (e.g., Meyer & Kump, 2008; Geochimica et Cosmochimica Acta , 1396 1414. Buhring€ SI, Sievert SM, Jonkers HM, Ertefai T, Elshahed MS, Johnston et al., 2009; Canfield, 2013; Leavitt et al., Krumholz LR, Hinrichs K-U (2011) Insights into 2013). chemotaxonomic composition and carbon cycling of phototrophic communities in an artesian sulfur-rich spring (Zodletone, Oklahoma, USA), a possible analog for ancient ACKNOWLEDGMENTS microbial mat systems. Geobiology 9, 166–179. Campbell BJ, Engel AS, Porter ML, Takai K (2006) The versatile This project was funded by the National Science Foundation ɛ-proteobacteria: key players in sulphidic habitats. Nature (NSF EAR-0525503 to A.P. and J.L.M.) and the Gordon Reviews Microbiology 4, 458–468. and Betty Moore Foundation (to A.P.). T.L.H. graciously Canfield DE (1998) A new model for Proterozoic ocean acknowledges support from the NAI Postdoctoral Program. chemistry. Nature 396, 450–453. W.M. was supported by a Marie Curie International Outgo- Canfield DE (2013) Sulfur isotopes in coal constrain the evolution of the Phanerozoic sulfur cycle. Proceedings of the National ing Fellowship within the 7th European Community Frame- Academy of Science 110, 8443–8446. work Programme. Funds for sample collection and the Canfield DE, Stewart FJ, Thamdrup B, De Brabandere L, complementary geochemical context were provided by the Dalsgaard T, Delong EF, Revsbech NP, Ulloa O (2010) A NASA Exobiology Program (T.W.L.). Work in the labora- cryptic sulfur cycle in oxygen-minimum-zone waters off the 330 – tory of D.A.B. on genomes of chlorophototrophic organ- Chilean coast. Science , 1375 1378. Caumette P, Schmidt K, Biebl H, Pfennig N (1985) isms was supported by the National Science Foundation Characterization of a Thiocapsa strain containing okenone as (MCB-0523100) and NASA (NNX09AM87G). major carotenoid. Systematic and Applied Microbiology 6, 132– 136. Caumette P, Baulaigue R, Matheron R (1991) Thiocapsa halophila REFERENCES sp. nov., a new halophilic phototrophic sulfur strain. Archives of Microbiology 155, 170–176. Aller RC (1998) Mobile deltaic and continental shelf muds as Caumette P, Guyoneaud R, Imhoff JF, Suling€ J, Gorlenko V 61 – suboxic, fluidized bed reactors. Marine Chemistry , 143 (2004) Thiocapsa marina sp. nov., a novel, okenone-containing, 155. purple sulfur bacterium isolated from brackish coastal and Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ (1990) marine environments. International Journal of Systematic and Basic local alignment search tool. Journal of Molecular Biology Evolutionary Microbiology 54, 1031–1036. 215 – , 403 410. Coolen MJL, Overmann J (1998) Analysis of subfossil molecular Amend JP, Shock ES (2001) Energetics of overall metabolic remains of purple sulfur bacteria in a lake sediment. Applied and reactions of thermophilic and hyperthermophilic Archaea and Environmental Microbiology 64, 4513–4521. 25 – Bacteria. FEMS Microbiology Reviews , 175 243. Dahl C, Schulte A, Stockdreher Y, Hong C, Grimm F, Sander J, Anisimova M, Gascuel O (2006) Approximate likelihood-ratio test Kim R, Kim S-H, Shin DH (2008) Structural and molecular for branches: a fast, accurate, and powerful alternative. Systems genetic insight into a wide-spread bacterial sulfur oxidation 55 – Biology , 539 552. pathway. Journal of Molecular Biology 384, 1287–1300. Aziz RK, Bartels D, Best AA, DeJongh M, Disz T, Edwards RA, Dahl C, Franz B, Hensen D, Kesselheim A, Zigann R (2013) Formsma K, Gerdes S, Glass EM, Kubal M, Meyer F, Olsen GJ, Sulfite oxidation in the purple sulfur bacterium Allochromatium Olson R, Osterman AL, Overbeek RA, McNeil LK, Paarmann vinosum: identification of SoeABC as a major player and D, Paczian T, Parrello B, Pusch GD, Reich C, Stevens R, relevance of SoxYZ in the process. Microbiology 159, 2626– Vassieva O, Vonstein V, Wilke A, Zagnitko O (2008) The 2638. RAST Server: rapid annotations using subsystems technology. Darriba D, Taboada G, Doallo R, Posada D (2012) jModelTest 2: 9 BMC Genomics , 75. more models, new heuristics and parallel computing. Nature Baumgartner LK, Reid RP, Dupraz C, Decho AW, Buckley DH, Methods 9, 772. Spear JR, Przekop KM, Visscher PT (2006) Sulfate reducing Decristophiris PMA, Peduzzi S, Ruggeri-Bernardi N, Hahn D, bacteria in microbial mats: changing paradigms, new discoveries. Tonolla M (2009) Fine scale analysis of shifts in bacterial 185 – Sedimentary Geology , 131 145. community structure in the chemocline of meromictic Lake € Biegel E, Muller V, Schmidt S, Gonzalez JM (2011) Cadagno, Switzerland. Journal of Limnology 68,16–24. Biochemistry, evolution and physiological function of the Rnf

© 2014 John Wiley & Sons Ltd 466 T. L. HAMILTON et al.

Del Don C, Hanselmann KW, Peduzzi R, Bachofen R (2001) The Hedderich R, Klimmek O, Kroger€ A, Dirmeier R, Keller M, meromictic alpine Lake Cadagno: orographical and Stetter KO (1998) with elemental biogeochemical description. Aquatic Sciences 63,70–90. sulfur and with disulfides. FEMS Microbiology Reviews 22, Dick JD, Andersson AF, Baker BJ, Simmons SL, Thomas BC, 353–381. Yelton AP, Banfield JF (2009) Community-wide analysis of Hensen D, Sperling D, Truper€ HG, Brune DC, Dahl C (2006) microbial genome sequence signatures. Genome Biology 10, Thiosulphate oxidation in the phototrophic sulphur R85. strain Allochromatium vinosum. Molecular Microbiology 62, Finster K (2008) Microbiological disproportionation of inorganic 794–810. sulfur compounds. Journal of Sulfur Chemistry 29, 281–292. Holkenbrink C, Ocon Barbas S, Mellerup A, Otaki H, Frigaard Frigaard NU, Bryant DA (2008) Genomic insights into the sulfur N-U (2011) Sulfur globule oxidation in green sulfur bacteria is metabolism of phototrophic green sulfur bacteria. In Sulfur dependent on the dissimilatory sulfite reductase system. Metabolism in Phototrophic Organisms (eds Hell R, Dahl C, Knaff Microbiology 157, 1229–1239. DB, Leustak T). Springer Science, New York, pp. 337–355. Hug LA, Castelle CJ, Wrighton KC, Thomas BC, Sharon I, Frigaard NU, Dahl C (2009) Sulfur metabolism in phototrophic Frischkorn KR, Williams KH, Tringe SG, Banfield JF (2013) sulfur bacteria. Advances in Microbial Physiology 54, 103–200. Community genomic analyses constrain the distribution of Ghosh W, Dam B (2009) Biochemistry and molecular biology of metabolic traits across the Chloroflexi phylum and indicate roles lithotrophic sulfur oxidation by taxonomically and ecologically in sediment carbon cycling. Microbiome 1, 22. diverse bacteria and archaea. FEMS Microbiology Reviews 33, Huson DH, Mitra S, Weber N, Ruscheweyh H, Schuster SC 999–1043. (2011) Integrative analysis of environmental sequences using Gregersen LH, Habicht KS, Peduzzi S, Tonolla M, Canfield DE, MEGAN4. Genome Research 21, 1552–1560. Miller M, Cox RP, Frigaard N-U (2009) Dominance of a clonal Imhoff JF (2001) Transfer of Pfennigia purpurea Tindall 1999 green sulfur bacterial population in a stratified lake. FEMS (Amoebobacter purpureus Eichler and Pfennig 1988) to the Microbiology Ecology 70,30–41. genus Lamprocystis as Lamprocystis purpurea comb. nov. Gregersen LH, Bryant DA, Frigaard N-U (2011) Mechanisms and International Journal of Systematic and Evolutionary evolution of oxidative sulfur metabolism in green sulfur bacteria. Microbiology 51, 1699–1701. Frontiers in Microbiology 2, 116. Imhoff JF (2005) Order I. Chromatiales ord. nov. In Bergey’s Grein F, Pereira IAC, Dahl C (2010) Biochemical characterization Manual of Systematic Bacteriology (eds Brenner DJ, Krieg NR, of individual components of the Allochromatium vinosum Stanley JT), 2nd ed., vol 2., part B. The Gammaproteobacteria. DsrMKJOP transmembrane complex aids understanding of Springer, New York, pp. 1–3. complex function in vivo. Journal of Bacteriology 24, 6369–6377. Inagaki F, Takai K, Kobayashi H, Nealson KH, Horikoshi K Grein F, Ramos AR, Venceslau SS, Pereira IAC (2013) Unifying (2003) Sulfurimonas autotrophica gen. nov., sp. nov., a novel concepts in anaerobic respiration: insights from dissimilatory sulfur-oxidizing e-proteostrain isolated from hydrothermal sulfur metabolism. Biochimica et Biophysica Acta 1827, 145–160. sediments in the Mid-Okinawa Trough. International Journal of Grice K, Cao CQ, Love GD, Bottcher ME, Twitchett RJ, Systematic and Evolutionary Microbiology 53, 1801–1805. Grosjean E, Summons RE, Turgeon SC, Dunning W, Jin YG Iverson V, Morris RM, Frazar CD, Berthiaume CT, Morales RL, (2005) Photic zone euxinia during the Permian-Triassic Armbrust EV (2012) Untangling genomes from metagenomes: superanoxic event. Science 307, 706–709. revealing an uncultured class of marine Euryarchaeota. Science Grimm F, Franz B, Dahl C (2008) Thiosulfate and sulfur 335, 587–590. oxidation in purple sulfur bacteria. In Microbial Sulfur Jeong HS, Jouanneau Y (2000) Enhanced nitrogenase activity in Metabolism (eds Dahl C, Friedrich CG). Springer-Verlag, strains of Rhodobacter capsulatus that overexpress the rnf genes. Heidelberg, pp. 101–116. Journal of Bacteriology 182, 1208–1214. Guyoneaud R, Suling€ J, Petri R, Matheron R, Caumette P, Johnston DT, Wolfe-Simon F, Pearson A, Knoll AH (2009) Pfennig N, Imhoff JF (1998) Taxonomic rearrangements of the Anoxygenic photosynthesis modulated Proterozoic oxygen and genera Thiocapsa and Amoebobacter on the basis of 16S rDNA sustained Earth’s middle age. Proceedings of the National sequence analyses, and description of Thiolamprovum gen. nov. Academy of Sciences of the United States of America 106, International Journal of Systematic and Evolutionary Biology 48, 16925–16929. 957–964. Jones DS, Albrecht HL, Dawson KS, Schaperdoth I, Freeman Hadas O, Pinkas R, Erez J (2001) High chemoautotrophic KH, Pi Y, Pearson A, Macalady JL (2012) Community genomic primary production in Lake Kinneret, Israel: a neglected link in analysis of an extremely acidophilic sulfur-oxidizing biofilm. the carbon cycle of the lake. Limnology and Oceanography 46, ISME Journal 6, 158–170. 1968–1976. Jorgensen BB, Cohen Y, Revsbech NP (1986) Transition from Hall KJ, Northcote TG (1990) Production and decomposition anoxygenic to oxygenic photosynthesis in a microcoleus processes in a saline meromictic lake. Hydrobiologia 197, 115– chthonoplastes cyanobacterial mat. Applied and Environmental 128. Microbiology 51, 408–417. Handley KM, VerBerkmoes NC, Steefel CI, Williams KH, Sharon Klepac-Ceraj V, Hayes CA, Gilhooly WP, Lyons TW, Kolter R, I, Miller C, Frischkorn KR, Chourey K, Thomas BC, Shah MB, Pearson A (2012) Microbial diversity under extreme euxinia: Long PE, Hettich RL, Banfield JF (2012) Biostimulation Mahoney Lake, Canada. Geobiology 10, 223–235. induces syntrophic interactions that impact C, S, and N Kumar PA, Srinivas TNR, Sasikala Ch, Ramana ChV (2007) cycling in a sediment microbial community. ISME Journal 7, Halochromatium roseum sp. nov., a non-motile phototrophic 800–816. gammaproteobacterium with gas vesicles, and amended Hanson TE, Tabita FR (2003) Insights into the stress response description of the genus Halochromatium. International Journal and sulfur metabolism revealed by proteome analysis of a of Systematic and Evolutionary Microbiology 57, 2110–2113. Chlorobium tepidum mutant lacking the Rubisco-like protein. Leavitt WD, Halevy I, Bradley AS, Johnston DT (2013) Influence Photosynthetic Research 78, 231–248. of sulfate reduction rates on the Phanerozoic sulfur isotope

© 2014 John Wiley & Sons Ltd Sulfur cycle in a meromictic lake 467

record. Proceedings of the National Academy of Sciences of the Overmann J, Beatty JT, Hall KJ (1996b) Purple sulfur bacteria United States of America 110, 11244–11249. control the growth of aerobic heterotrophic bacterioplankton in Li H, Durbin R (2009) Fast and accurate short read alignment a meromictic salt lake. Applied and Environmental Microbiology with Burrows-Wheeler Transform. Bioinformatics 25, 62, 3251–3258. 1754–1760. Overmann J (1997) Mahoney Lake: a case study of the ecological Lohse M, Bolger AM, Nagel A, Fernie AR, Lunn JE, Stitt M, significance of phototrophic purple sulfur bacteria. In Advances Usadel B (2012) RobiNA: a user-friendly, integrated software in Microbial Ecology (ed. Jones JG). Plenum Press, New York, solution for RNA-Seq-based transcriptomics. Nucleic Acids pp. 251–288. Research 40, W622–W627. Pancost RD, Crawford N, Magness S, Turner A, Jenkyns HC, Meuser JE, Baxter BK, Spear JR, Peters JW, Posewitz MC, Boyd Maxwell JR (2004) Further evidence for the development of ES (2013) Contrasting patterns of community assembly in the photic-zone euxinic conditions during Mesozoic stratified water column of Great Salt Lake, Utah. Microbial oceanic anoxic events. Journal of the Geological Society 161, Ecology 66, 268–280. 353–364. Meyer KM, Kump LR (2008) Oceanic euxinia in Earth history: Peduzzi S, Storelli N, Weslh A, Peduzzi R, Hahd D, Perret X, causes and consequences. Annual Review of Earth and Tonolla M (2012) Candidatus “Thiodictyon syntrophicum”, sp. Planetary Sciences 36, 251–288. nov., a new purple sulfur strain isolated from the chemocline of Meyer KM, Macalady JL, Fulton JM, Kump LR, Schaperdoth I, Lake Cadagno forming aggregates and specific associations with Freeman KH (2011) Carotenoid biomarkers as an imperfect Desulfocapsa sp. Systematic and Applied Microbiology 35, 139– reflection of the anoxygenic phototrophic community in 144. meromictic Fayetteville Green Lake. Geobiology 9, 321–329. Peng Y, Leung HCM, Yiu SM, Chin FYL (2012) IDBA-UD: a de Miller CS, Baker BJ, Thomas BC, Singer SW, Banfield JF (2011) novo assembler for single-cell and metagenomic sequencing data EMIRGE: reconstruction of full-length ribosomal genes from with highly uneven depth. Bioinformatics 28, R69. microbial community short read sequencing data. Genome Pereira IAC, Ramos AR, Grein F, Marques MC, da Silva MS, Biology 12, R44. Venceslau S (2011) A comparative genomic analysis of energy Northcote TG, Hall KJ (1983) Limnological contrasts and metabolism in sulfate reducing bacteria and archaea. Frontiers in anomalies in 2 adjacent saline lakes. Hydrobiologia 105, 179–194. Microbiology 2,1. Northcote TG, Halsey TG (1969) Seasonal changes in the Pitcher RS, Watmough NJ (2004) The bacterial cytochrome cbb3 limnology of some meromictic lakes in southern British oxidases. Biochimica et Biophysica Acta 1655, 388–399. Columbia. Journal of Fisheries Research Board of Canada 26, Rabus R, Hansen TA, Widdel F (2013) Dissimilatory sulfate- and 1763–1787. sulfur-reducing prokaryotes. In The Prokaryotes – Prokaryotic Ogawa T, Furusawa T, Nomura R, Seo D, Hosoya-Matsuda N, Physiology and Biochemistry (eds Rosenbrug E, DeLong EF, Sakurai H, Inoue K (2008) SoxAX binding protein, a novel Lory S, Stackebrandt E, Thompson F). Springer-Verlag, component of the thiosulfate-oxidizing multienzyme system in Heidelberg, pp. 309–404. the green sulfur bacterium Chlorobium tepidum. Journal of Raes J, Korbel JO, Lercher MJ, von Mering C, Bork P (2007) Bacteriology 190, 6097–6110. Prediction of effective genome size in metagenomic samples. Oliveira TF, Vonrhein C, Matia PM, Venceslau SS, Pereira IAC, Genome Biology 8, R10. Archer M (2008) The crystal structure of Desulfovibrio vulgaris Reinhard CT, Planavsky NJ, Robbins LJ, Partin CA, Gill BC, dissimilatory sulfite reductase bound to DsrC provides novel Lalonde SV, Bekker A, Konhauser KO, Lyons TW (2013) insights into the mechanism of sulfate respiration. Journal of Proterozoic ocean redox and biogeochemical stasis. Proceedings Biological Chemistry 283, 34141–34149. of the National Academy of Sciences of the United States of Overmann J, Pfennig N (1989) Pelodictyon phaeoclathratiforme sp. America 110, 5357–5362. nov., a new brown-colored member of the Chlorobiaceae forming Sander J, Engels-Schwarzlose S, Dahl C (2006) Importance of the net-like colonies. Archives of Microbiology 152, 401–406. DsrMKJOP complex for sulfur oxidation in Allochromatium Overmann J, Beatty JT, Hall KJ, Pfennig N, Northcote TG vinosum and phylogenetic analysis of related complexes in other (1991) Characterization of a dense, purple sulfur bacterial layer prokaryotes. Archives of Microbiology 186, 357–366. in a meromictic salt lake. Limnology and Oceanography 36, 846– Schauder R, Preub A, Jetten M, Fuchs G (1989) Oxidative and 859. reductive acetyl CoA/carbon monoxide dehydrogenase pathway Overmann J, Fischer U, Pfennig N (1992) A new purple sulfur in Desulfostrain autotrophicum. Archives of Microbiology 151, bacterium from saline littoral sediments, Thiorhodovibrio 84–89. winogradskyi gen. nov. and sp. nov. Archives of Microbioogy 157, Shah N, Tang H, Doak TG, Ye Y (2010) Comparing bacterial 329–335. communities inferred from 16S rRNA gene sequencing and Overmann J, Pfennig N (1992) Bouyancy regulation and shotgun metagenomics. Biocomputing Proceedings of the aggregate formation in Amoebobacter purpureus from Mahoney Pacific Symposium pp. 165–176. Lake. FEMS Microbiology Letters 101,67–79. Simon J, Klotz MG (2013) Diversity and evolution of Overmann J, Sandmann G, Hall KJ, Northcote TG (1993) Fossil bioenergetic systems involved in microbial nitrogen and paleolimnology of meromictic Mahoney Lake, compound transformations. Biochimica et Biophysica Acta British Columbia, Canada. Aquatic Sciences 55,31–39. 1827, 114–135. Overmann J, Beatty JT, Hall KJ (1994) Photosynthetic activity Sorek R, Zhu Y, Creevey CJ, Francino MP, Bork P, Rubin EM and population dynamics of Amoebobacter purpureus in (2007) Genome-wide experimental determination of barriers to a meromictic saline lake. FEMS Microbiology Ecology 15, . Science 318, 1449–1452. 309–320. Storelli N, Peduzzi S, Saad MM, Frigaard N-U, Perret X, Tonolla

Overmann J, Beatty JT, Krouse HR, Hall KJ (1996a) The sulfur M (2013) CO2 assimilation in the chemocline of Lake Cadagno cycle in the chemocline of a meromictic salt lake. Limnology and is dominated by a few types of phototrophic purple sulfur Oceanography 41, 147–156. bacteria. FEMS Microbiology Ecology 84, 421–432.

© 2014 John Wiley & Sons Ltd 468 T. L. HAMILTON et al.

Swofford DL (2001) Paup: Phylogenetic Analysis Using Parsimony Table S3 Marker genes used to evaluate genome completion of the Delta- (and Other Methods), 4.0b10 edn. Sinauer Associate, proteobacteria community genome bins. Sunderland, MA. Tabita FR, Hanson TE, Li H, Statgopan S, Singh J, Chan S Table S4 Orthologous group markers and their abundance used to estimate (2007) Function, structure, and evolution of the RubisCO-like the number of complete genomes in the Deltaproteobacteria community proteins and their RubisCO homologs. Microbiology and genome bins. Molecular Biology Reviews 71, 576–599. Table S5 Single-copy phylogenetic marker genes used for taxonomic Tank M, Thiel V, Imhoff JE (2009) Phylogenetic relationship of assignment of the ML-Gamma and the ML-Epsilon genomic bins. phototrophic purple sulfur bacteria according to pufL and pufM 12 – genes. International Microbiology , 175 185. Table S6 List of genes from the Mahoney Lake 7-m metagenome identified Tuschak C, Leung MM, Beatty JT, Overmann J (2005) The puf and included in functional gene phylogenetic analyses. operon of the purple sulfur bacterium Amoebobacter purpureus, transcription and phylogenetic analysis. Archives of Microbiology Table S7 Number and taxonomic affiliation of key genes involved in the 138, 431–443. sulfur cycle. Tyson GW, Chapman J, Hugenholtz P, Allen EE, Ram RJ, Richardson PM, Solovyev VV, Rubin EM, Rokhsar DS, Banfield Table S8 Accession numbers of the Sqr sequences included in Fig. S6. JF (2004) Community structure and metabolism through Table S9 Summary of community genomic bins of Deltaproteobacteria spp. reconstruction of microbial genomes from the environment. 428 – Nature ,37 43. Table S10 Calculations for DG = DG0 + RTlnQ at 18°C for selected sulfur Vogl K, Bryant DA (2011) Elucidation of the biosynthetic redox reactions in situ. pathway for okenone in Thiodictyon sp. CAD16 leads to the discovery of two novel carotene ketolases. Journal of Biological Table S11 List of organisms from which NifH/VnfH/AnfH were identified Chemistry 286, 38521–38532. and included in the phylogenetic analysis of NifH/VnfH/AnfH identified in Vogl K, Bryant DA (2012) Biosynthesis of the biomarker from the Mahoney Lake 7-m metagenome. okenone: v-ring formation. Geobiology 10, 205–215. Voorhies AA, Biddanda BA, Kendall T, Jain S, Marcus DN, Nold Fig. S1 ESOM (repeating-tiled view) bins of the tetranucleotide frequency signatures of the 7-m metagenome with bins of interest indicated. SC, Sheldon ND, Dick GJ (2012) Cyanobacterial life at low O2: community genomics and function reveal metabolic versatility Fig. S2 The estimated number of SRB genomes recovered from the Mahon- and extremely low diversity in a Great Lakes sinkhole mat. ey Lake metagenome. Geobiology 10, 250–267. Whitaker R, Banfield J (2006) Population genomics in natural Fig. S3 Maximum likelihood phylogenetic tree of 19 concatenated single- 21 microbial communities. Trends in Ecology and Evolution , copy ribosomal proteins showing the placement of the strain ML1 and ML- – 508 516. Gamma compared to closely related Gammaproteobacteria spp. for which Wrighton KC, Thomas BC, Sharon I, Miller CS, Castelle CJ, genome sequences are available. VerBerkmoes NC, Wilkins MJ, Hettich RL, Lipton MS, Williams KH, Long PE, Banfield JF (2012) Fermentation, Fig. S4 Phylogenetic diversity of RuBisCO in the Mahoney Lake chemo- hydrogen, and sulfur metabolism in multiple uncultivated cline. Maximum likelihood phylogenetic trees constructed from CbbL and . Science 337, 1661–1665. CbbM sequences mined from the ML 7-m metagenome and the NCBI and Wu M, Eisen JA (2008) A simple, fast, and accurate method of IMG/M database. phylogenomic inference. Genome Biology 9, R151. Zerbino DR, Birney E (2008) Velvet: algorithms for de novo Fig. S5 Phylogenetic diversity of the nitrogenase Fe protein in the Mahoney short read assembly using de Bruijn graphs. Genome Research Lake chemocline. 18, 821–829. Fig. S6 Maximum likelihood phylogenetic tree of translated genes Zerkle AL, Kamyshny A Jr, Kump LR, Farquhar J, Oduro H, sequences encoding okenone C-4/4’ ketolase (CruO), okenone accessory Arthur MA (2010) Sulfur cycling in a stratified euxinic lake with ketolase (CruS), carotenoid desaturase (CrtD) and methoxyneurosporine de- moderately high sulfate: constraints from quadruple S isotopes. hydrogenases (CrtD/CrtI family homologs). Geochimica et Cosmochimica Acta 74, 4953–4970. Zopfi J, Ferdelman TG, Fossing G (2004) Distribution and fate of Fig. S7 Maximum likelihood phylogenetic tree of Sqr. Sequences were mined – sulfur intermediates sulfite, tetrathionite, thiosulfate, and from the 7-m metagenome (Table S6) and the NCBI and IMG/M database. elemental sulfur–in marine sediments. In Sulfur Biogeochemistry: Past and Present (eds Amend JP, Edwards KJ, Lyons TW). The Fig. S8 Taxonomic diversity of SRB-affiliated SP3 proteins and the Predicted Geological Society of America, Boulder, CO, pp. 97–117. role of the ML SRB spp. in the sulfur cycle.

Fig. S9 Phylogenetic diversity of RuBisCO in the Mahoney Lake chemocline. SUPPORTING INFORMATION Maximum likelihood phylogenetic trees constructed from CbbL and CbbM sequences mined from the ML 7-m metagenome and the NCBI and IMG/M Additional Supporting Information may be found in the database. online version of this article: Fig. S10 Calculated values for DG = DG0 + RTlnQ at 18 °C for selected sul- Data S1 Supplementary methods and results. fur redox reactions in situ, at depths above, within, and below the purple chemocline. Table S1 Metagenome summary statistics.

Table S2 Marker genes used to evaluate genome completion of the isolate genome, strain ML1, and the ML-Gamma and ML-Epsilon genomic bins.

© 2014 John Wiley & Sons Ltd