Abstract
Microorganisms have driven Earth’s sulfur cycle since the emergence of life1,2,3,4,5,6, yet the sulfur-cycling capacities of microorganisms and their integration with other element cycles remain incompletely understood. One such uncharacterized metabolism is the coupling of sulfide oxidation with iron(iii) oxide reduction, a ubiquitous environmental process hitherto considered to be strictly abiotic7,8. Here we present a comprehensive genomic analysis of sulfur metabolism across prokaryotes, and reveal bacteria that are capable of oxidizing sulfide using extracellular solid phase iron(iii). Based on a phylogenetic framework of over hundred genes involved in dissimilatory transformation of sulfur compounds, we recorded sulfur-cycling capacity in most bacterial and archaeal phyla. Metabolic reconstructions predicted co-occurrence of sulfur compound oxidation and iron(iii) oxide respiration in diverse members of 37 prokaryotic phyla. Physiological and transcriptomic evidence demonstrated that a cultivated representative, Desulfurivibrio alkaliphilus, grows autotrophically by oxidizing dissolved sulfide or iron monosulfide (FeS) to sulfate with ferrihydrite as an extracellular iron(iii) electron acceptor. The biological process outpaced the abiotic process at environmentally relevant sulfide concentrations. These findings expand the known diversity of sulfur-cycling microorganisms and unveil a biological mechanism that links sulfur and iron cycling in anoxic environments, thus highlighting the fundamental role of microorganisms in global element cycles.
Similar content being viewed by others
Main
The global biogeochemical sulfur cycle is largely driven by bacteria and archaea that have evolved an array of enzymatic mechanisms for diverse dissimilatory sulfur redox transformations (hereafter called sulfur microorganisms). Sulfur transformations are linked to redox reactions of other elements, such as carbon, oxygen, iron and nitrogen1,2,3. The intertwined electron transfer reactions mediated by sulfur microorganisms in diverse ecosystems represent a metabolic network on the global scale that profoundly impacts biogeochemistry, climate and redox state on the Earth’s surface over geological timescales4,5,6. Functional gene and genome-centric surveys have expanded the known diversity of specific guilds of sulfur microorganisms beyond cultivated representatives, and predicted previously unrecognized roles of uncultivated sulfur microorganisms in biogeochemical cycles across ecosystems9,10,11. However, the microbial genomic signatures for the full spectrum of dissimilatory sulfur metabolisms, including oxidation of reduced sulfur compounds (hereafter called sulfur oxidation), has not been explored. Moreover, most genome-based predictions of sulfur-related metabolisms lack support from experimental evidence.
Among the various transformation processes catalysed by sulfur microorganisms, the re-oxidation of sulfide, which occurs mainly by dissimilatory reduction of oxidized sulfur compounds such as sulfate, is a major biogeochemical process. Sulfur oxidation replenishes sulfate and contributes to the sulfur cycle in diverse ecosystems12,13,14. Microorganisms oxidize reduced sulfur compounds by utilizing light or chemical oxidants, including oxygen, nitrate and manganese oxides15. Ferric oxyhydroxides and oxides (hereafter called iron(iii) oxides) represent one of the largest pools of oxidants on the Earth’s surface16. The interaction with iron(iii) oxides shapes the concentration of free sulfide and the cycling of sulfur in various anoxic environments, such as marine sediments, wetlands and aquifers7,12,17,18,19. However, current biogeochemical models consider the reaction of sulfide with iron(iii) oxides as purely abiotic, producing mainly elemental sulfur (S(0)) and poorly crystalline FeS7,8. Although geochemical studies have hinted at microbial oxidation of sulfide to sulfate with iron(iii) oxides20,21,22, microorganisms and reactions that catalyse this process thus far remained uncharacterized.
Here we report the wide distribution of dissimilatory sulfur-transforming potential across the bacterial and archaeal tree of life, including many previously unsuspected microbial taxa. By genome-based reconstruction of the metabolism of potential sulfur-oxidizing microorganisms, we further revealed sulfur-based electron transfer pathways that could facilitate anaerobic sulfur oxidation coupled to the reduction of extracellular iron(iii) oxide in diverse bacterial and archaeal taxa. Physiological and transcriptomic evidence of the predicted process in a cultivated representative, D. alkaliphilus, contributes to the evolving view on the biological coupling of the biogeochemical sulfur and iron cycles.
Sulfur-metabolizing bacteria and archaea
We first established a computational framework to more accurately predict dissimilatory sulfur metabolism from bacterial and archaeal genomes (Supplementary Fig. 1). To this end, we performed phylogenetic analyses of 116 proteins that are involved in diverse sulfur redox transformations, and developed hidden Markov models (HMMs) for all monophyletic clades that corresponded to functional homologues of experimentally validated proteins (Supplementary Text). The resulting phylogenies and clade-specific HMMs enabled us to effectively and accurately retrieve homologues of sulfur-cycling proteins from genomes and curate their function within a phylogenetic context (Supplementary Fig. 2).
Using this resource, we systematically queried a subset of 42 sulfur-cycling enzymes against representative genomes of all bacterial and archaeal species from the Genome Taxonomy Database (GTDB)23. The selected enzymes are key markers for specific, mostly dissimilatory sulfur metabolisms (Supplementary Table 1). More than half of the species, representing 120 (80.5%) of 149 known bacterial and archaeal phyla, encoded at least one sulfur-cycling marker protein (Fig. 1). Most genomes that encode a marker protein (for example, DsrA) also encode additional proteins (for example, DsrB, DsrC, DsrE and DsrF) associated with the respective marker protein for coordinated enzymatic function (Supplementary Fig. 3). The occurrence of sulfur-cycling capability in most bacterial and archaeal phyla reflects the deep integration of sulfur redox processes into microbial metabolism since the origin of life2,5,6. In this context, 5,561 species with sulfur-cycling potential, affiliated to 71 phyla, were represented exclusively by genomes from so far uncultured microorganisms (Fig. 1). This highlights the broad taxonomic distribution of sulfur-cycling potential and underscores that many of the microorganisms that we identified here as capable of sulfur-cycling remain uncultured and poorly characterized. Our overview of putative sulfur microorganisms provides a foundation for more detailed metabolic predictions and experimental validation.
a, Left, number of GTDB taxa carrying at least one of 42 sulfur-cycling marker genes. Right, among sulfur-metabolizing microorganisms, a substantial fraction is exclusively represented by uncultured taxa. b, Distribution of 42 sulfur-cycling marker genes across archaeal and bacterial phyla. The top 50 of 120 phyla with the highest number of sulfur-cycling genes are shown. The size of points indicates the number of genomes encoding a particular sulfur-cycling gene (with log transformation). Sulfur-cycling guilds lacking any cultivated representatives in a phylum are depicted in less opaque colour. Double asterisks indicate microbial phyla with no or few cultivated representatives. Genes associated with broad sulfur-cycling functional categories are shown with different colours. DMS, dimethylsulfide; SQ, sulfoquinovose.
Microbial S oxidation with iron(iii) oxide
We next screened the genomes of the identified sulfur microorganisms for potential electron transfer metabolisms that facilitate anaerobic sulfur oxidation by reducing iron(iii) oxides, a microbial physiology of potentially broad biogeochemical importance in marine and terrestrial ecosystems24. In contrast to coupling of elemental, zero-valent sulfur oxidation to dissolved iron(iii) reduction by acidophiles in highly acidic environments25,26 (pH < 3), no microorganism has been shown to gain energy from coupled reduction of solid phase extracellular iron(iii) oxides and oxidation of reduced sulfur compounds. We identified co-occurring genetic features for dissimilatory sulfur oxidation and extracellular iron(iii) reduction metabolisms across 37 bacterial and archaeal phyla (Supplementary Text and Supplementary Figs. 4 and 5). Reconstruction of sulfur and iron energy conservation pathways revealed three metabolic options (Fig. 2). The first option couples iron(iii) reduction to the oxidation of sulfide to sulfate (reaction 1). The underlying metabolic pathway was found in, for example, members of Desulfurivibrionaceae, whose genomes encode a full suite of enzymes (sulfate adenylyltransferase (Sat), adenosine-5′-phosphosulfate reductase (AprAB) and dissimilatory sulfite reductase (DsrAB)) for the canonical sulfate reduction pathway (Supplementary Fig. 5). The DsrAB sequences of Desulfurivibrionaceae form a well-supported monophyletic clade, including D. alkaliphilus, which operates DsrAB in reverse during sulfide oxidation to sulfate27 (Supplementary Fig. 6). The same genomes also contain multiple homologues of Geobacter-type cytochromes, which are necessary for dissimilatory iron(iii) reduction (Supplementary Fig. 5), including extracellular OmcS and porin–cytochrome complex28,29,30 (OmaB–OmbB–OmcB). The interactions of these enzymes would establish a conduit channeling sulfide-derived electrons to insoluble extracellular substrates such as iron(iii) oxides. The second option involves a sulfide:quinone oxidoreductase (Sqr) and FccBA for sulfide oxidation, and a multi-haem protein complex (MtrCAB) for extracellular respiration of iron(iii) oxides (Supplementary Fig. 7). Homologues of Sqr, FccBA and MtrCAB were detected in two uncultured Rhodoferax species. The Sqr is affiliated to type I Sqr that has a physiological role in sulfide-based energy transduction, as described in Aquifex aeolicus31 and Rhodobacter capsulatus32. The MtrCAB homologue encoded in the Rhodoferax genome is closely related to MtrCAB from the known iron(iii) reducer Rhodoferax ferrireducens33,34. These genetic systems could support an energy metabolism that couples iron reduction with dissimilatory oxidation of sulfide to elemental sulfur (reaction 2). Additionally, MtrCAB was detected in multiple known thiosulfate oxidizers (Supplementary Fig. 5), including cultivated members within Burkholderiaceae, Sulfurifustaceae, Thiohalomonadaceae and Ectothiorhodospiraceae. This suggests a third option of extracellular iron(iii)-dependent thiosulfate oxidation (reaction 3). Beyond Desulfobacterota and Proteobacteria, members from 35 other microbial phyla also have the potential to catalyse reactions 1–3 via different gene combinations (Supplementary Text and Supplementary Figs. 4 and 5).
Top, genomic reconstruction predicts that dissimilatory reduction of iron(iii) oxides is coupled with oxidation of sulfide to sulfate (reaction 1; a), oxidation of sulfide to elemental sulfur (reaction 2; b) and oxidation of thiosulfate to sulfate (reaction 3; c). Key enzymes from Desulfobacterota and Gammaproteobacteria taxa are indicated as examples. Bottom, Gibbs free energy (∆Gr) of the iron-dependent sulfur oxidation reactions across a range of pH was calculated, assuming two different environmental settings. Marine ecosystem conditions (red line): ionic strength (I) = 0.7 M, T = 25 °C, [total dissolved sulfide] = 100 μM, [SO42−] = 28 mM, [S2O32−] = 1 μM, [Fe2+] = 10 μM; freshwater ecosystem conditions (black line): I = 0.001 M, T = 25 °C, [total dissolved sulfide] = 100 μM, [SO42−] = 100 μM, [S2O32−] = 1 μM, [Fe2+] =10 μM. The red and black lines overlap in c. Sox, thiosulfate-oxidizing system; OmcS and OmcZ, outer membrane multi-haem c-type cytochromes; PCC, porin–cytochrome complex.
Calculation of the Gibbs free energy showed that all three reactions could provide energy to microorganisms, even in neutral to moderately alkaline conditions (Fig. 2). For example, iron(iii)-dependent sulfide oxidation under natural settings (that is, freshwater and marine sediment) yields −20 to −40 kJ per mole electron, which is considered sufficient to support microbial growth35. Overall, these findings suggest wider occurrence of sulfur oxidation-dependent energy metabolisms across microbial taxa than previously recognized and provide testable hypotheses regarding the physiology of uncultured and cultured microorganisms.
Physiology and transcriptome
We aimed to experimentally validate the genome-predicted potential to oxidize sulfide with iron(iii) oxide in D. alkaliphilus, which encodes the genomic repertoire for reaction 1. We prioritized validatation of this predicted sulfide metabolism given the pervasive co-occurrence of sulfide and iron(iii) oxides across a wide range of ecosystems7,18,36. D. alkaliphilus has a versatile sulfur metabolism. It grows by disproportionation of elemental sulfur or by coupling sulfide oxidation with nitrate reduction27,37 but was not known to reduce iron(iii) oxides. Here we expand the known physiology of D. alkaliphilus by showing its capability to reduce iron(iii) oxides (ferrihydrite) with formate, poorly crystalline FeS or dissolved sulfide as electron donor. Initial incubation of D. alkaliphilus with ferrihydrite and formate led to simultaneous formate consumption and Fe(ii) production, with the following stoichiometry: HCOO− + 2Fe(iii) → CO2 + 2Fe(ii) + H+. This verified its capacity to reduce extracellular solid iron(iii) oxides (Fig. 3a).
a, D. alkaliphilus oxidizes formate (triangles) while reducing ferrihydrite (Fh) to Fe(ii) (circles). No Fe(ii) or formate consumption occurred in the absence of Fh or formate. b, Left and middle, D. alkaliphilus couples oxidation of poorly crystalline FeS to sulfate with reduction of Fh. Sulfate and Fe(ii) formation were not detected in abiotic controls or in cultures with FeS or Fh alone. Right, the ratio of sulfate to Fe(ii) formation (solid line) is close to predicted 1:8 stoichiometry (dashed line). c, Sulfate, HCl-extractable Fe(ii), elemental sulfur and Cline-extractable sulfide during the incubation of D. alkaliphilus cultures with Fh and daily spike of sulfide (1 mM). Parallel incubations with killed cells, or living cells without Fh or sulfide served as controls. Arrows indicate sulfide spike. d, Kinetics of dissolved sulfide and first-order rate constant in the incubations described in c. e, The canonical biogeochemical scenario of sulfate formation considering chemical sulfide oxidation to elemental sulfur (S(0)), followed by bacterial disproportionation. f, Updated scenario considering the contribution of sulfate formation from microbial oxidation of sulfide or FeS with iron oxides. g, Biphasic sulfate formation by D. alkaliphilus during 24 h incubation with ferrihydrite and sulfide. Phases I and II differ by sulfide availability. Phase I is shown with a larger scale on the x axis for visualization. A control for phase II was conducted by inoculating cells after the chemical reaction between sulfide and ferrihydrite (cells + Fh + S(0)/FeS). h, Faster consumption of a low concentration (approximately 50 µM) of sulfide by D. alkaliphilus compared with chemical controls lacking cells. Three spikes of sulfide were supplied at 1.5 h intervals. Results with different cell density are shown in Extended Data Fig. 1. In a–d,g,h, triplicate cultures (n = 3) were used for each incubation condition.
We then fed D. alkaliphilus FeS as electron donor and ferrihydrite as electron acceptor (Fig. 3b). Over a five-day incubation period, we observed the concurrent production of sulfate and Fe(ii). By contrast, Fe(ii) production was not detected in autoclaved controls, indicating that FeS was not chemically oxidized by ferrihydrite to form more oxidized sulfur species38, such as S(0). This observation excludes the possibility of sulfate formation by microbial disproportionation of S(0). Sulfate formation by FeS oxidation with residual nitrate from the inoculum was also excluded, as addition of FeS alone (that is, without ferrihydrite addition) did not result in sulfate accumulation. The slight increase of sulfate in ferrihydrite-only controls is probably owing to residual sulfide from the inoculum. After accounting for sulfur transformations observed in control incubations, the calculated ratio of sulfate to Fe(ii) formation was close to the predicted stoichiometry of 1:8 for reaction 1 (Fig. 3b), indicating that D. alkaliphilus oxidizes FeS-sulfide to sulfate by reducing ferrihydrite. The slight deviation from the expected stoichiometry is likely to result from reverse electron flow for carbon fixation (Supplementary Text and Supplementary Figs. 8 and 9).
Finally, to examine whether D. alkaliphilus oxidizes free sulfide with iron(iii) as electron acceptor, we incubated the culture with ferrihydrite and supplied 1 mM dissolved sulfide daily (Fig. 3c). This led to progressive accumulation of sulfate, S(0), and Fe(ii). In sterile controls, S(0) and Fe(ii) accumulated, but not sulfate. Similarly, sulfate was not formed in biotic controls lacking either ferrihydrite or sulfide (Fig. 3c, Supplementary Fig. 10 and Supplementary Text). These results showed that sulfate was formed in a biological process requiring ferrihydrite. We reasoned that biological oxidation of dissolved sulfide with ferrihydrite contributed to sulfate formation, as evidenced by: (1) exclusion of sulfur disproportionation as the only source of sulfate; (2) a biphasic production of sulfate following the supply of dissolved sulfide; and (3) accelerated sulfide consumption with cells compared to abiotic controls. We consider these lines of physiological evidence separately in the following three paragraphs.
A plausible source of sulfate in our incubation is microbial disproportionation of S(0), derived from the chemical oxidation of sulfide by ferrihydrite12,39. Under this canonical biogeochemical scenario, consumption of dissolved sulfide is controlled by the availability of reactive sites on the surface of ferrihydrite7, and the rate should decrease over time with periodic supply of sulfide due to surface saturation and production of sulfide by S(0) disproportionation. However, tracing the kinetics of dissolved sulfide removal revealed accelerated consumption upon successive supply of 1 mM sulfide, whereas the sterile control followed the predicted trend (Fig. 3d,e). The faster rate indicates that additional processes consumed the dissolved sulfide. We suggest dissimilatory iron(iii) reduction fuelled by reduced sulfur species oxidation40 as an additional process that rapidly precipitates dissolved sulfide by producing excessive Fe(ii) (Fig. 3f). Depending on the sulfur speciation in our incubation, the sulfur compounds supporting biological iron(iii) reduction include dissolved sulfide, FeS and S(0). Their oxidation, together with S(0) disproportionation by D. alkaliphilus contributes to the observed sulfate formation.
To probe for direct oxidation of dissolved sulfide to sulfate by D. alkaliphilus, we measured sulfate formation upon sulfide addition (1 mM) to ferrihydrite at higher temporal resolution. We observed biphasic sulfate formation dynamics within 24 h (Fig. 3g). Within the first 70 min, when dissolved sulfide was available (phase I), sulfate was formed at an average rate of 0.69 fmol cell−1 h−1. However, the rate decreased by approximately tenfold (0.061 fmol cell−1 h−1) after the depletion of dissolved sulfide (phase II). To measure the effect of poorly soluble sulfur species S(0) and FeS on sulfate accumulation rates, we first allowed sulfide and ferrihydrite to react for 70 min, after which dissolved sulfide has been completely transformed into FeS and S(0), and only then inoculated cells. Here we also observed a low sulfate formation rate (0.047 fmol cell−1 h−1) comparable to the one in the sulfide-free phase II (Fig. 3g). These results demonstrated that the presence of poorly soluble sulfur species alone (such as S(0) and FeS) was insufficient to sustain the sulfate formation rate observed at phase I. We thus attribute the sulfide-dependent, high sulfate formation rate of D. alkaliphilus in phase I to direct biological sulfide oxidation with ferrihydrite.
To further show that D. alkaliphilus directly oxidizes dissolved sulfide with iron(iii) oxide, we tracked sulfide kinetics in the culture incubated with a low concentration of dissolved sulfide (approximately 50 µM) and ferrihydrite. We hypothesized that the low sulfide concentration reduces its chemical reaction rate with ferrihydrite, enabling a clear readout of the biological sulfide transformation process. Indeed, sulfide was readily consumed by the ferrihydrite-amended culture down to 10 µM within 25 min, whereas 20–30 µM of sulfide remained in controls lacking cells. Repetitive addition of sulfide showed the same recurring pattern. The culture rapidly consumed sulfide and its estimated rate constant of removal was significantly higher than of abiotic controls (P < 0.01, Student’s t-test; Extended Data Fig. 1). This pattern was observed reproducibly in independent incubations with different cell densities (Extended Data Fig. 1). By ruling out alternative biogeochemical processes that support accelerated sulfide consumption (Supplementary Text), we concluded that D. alkaliphilus participates in sulfide transformation using ferrihydrite. Repeated supply (n = 8) of sulfide showed the culture transformed more than 90% of spiked sulfide to sulfate with negligible S(0) formation, while producing an excessive amount of Fe(ii) (Extended Data Fig. 2). In the abiotic control, however, S(0)—but not sulfate—was the main oxidized sulfur compound and Fe(ii) accumulated to much lower levels (Extended Data Fig. 2). These results indicate that respiration of ferrihydrite with sulfide by D. alkaliphilus outpaces the chemical reaction rate between sulfide and ferrihydrite at environmentally relevant low sulfide concentrations.
To show that coupled sulfide and iron(iii) oxides metabolism supports growth, we monitored the cell density of D. alkaliphilus during 4- to 13-day incubation with ferrihydrite and sulfide. We observed twofold to threefold increases in cell numbers for ferrihydrite-incubated cultures periodically amended with either 1 mM dissolved sulfide, approximately 50 µM dissolved sulfide or FeS (Fig. 4a–c). In comparison, incubation of D. alkaliphilus with dissolved sulfide and nitrate led to a 5–6 fold increase of cell number over 3 days (Fig. 4d and Supplementary Text). The calculated specific growth rate was higher in ferrihydrite cultures with dissolved sulfide (0.288 ± 0.015 day−1 at 1 mM and 0.436 ± 0.074 day−1 at 50 µM) compared to those with solid phase FeS (0.089 ± 0.005 day−1) (Supplementary Fig. 11), probably owing to the limited solubility of FeS. By contrast, no or only minor increases in cell density were detected in the dissolved sulfide-only, FeS-only and ferrihydrite-only controls (Fig. 4a,b). These results demonstrate that D. alkaliphilus is capable of utilizing sulfide or FeS together with ferrihydrite for growth. We attribute the overall restricted growth with ferrihydrite to the low energy field and higher maintenance energy requirements at alkaline pH (Supplementary Text). To further explore the association between growth and carbon fixation, we supplemented parallel cultures with 13C-bicarbonate (10 atom%), ferrihydrite, and sulfide or FeS. Carbon isotope composition analysis of bulk biomass and single cells revealed enrichment of 13C in living cells after five-day incubation (Fig. 4e–i), reflecting a chemoautotrophic lifestyle of D. alkaliphilus when growing on ferrihydrite and dissolved sulfide or FeS.
a, Mean cell density (n = 3) increased significantly (P = 2.18 × 10−8; ANOVA) over 4 days of incubation with ferrihydrite and daily addition of sulfide (1 mM). b, Mean cell density (n = 3) increased significantly (P = 3.79 × 10−10; ANOVA) over 13 days of incubation with ferrihydrite and periodic spikes of FeS (approximately 1 mM). c, Mean cell density (n = 3) increased significantly (P = 1.84 × 10−8; ANOVA) over 5 days of incubation with ferrihydrite and periodic addition of approximately 50 µM sulfide. d, Mean cell density (n = 3) increases significantly (P = 5.53 × 10−11; ANOVA) alongside sulfide consumption over 3 days of incubation with sulfide and nitrate (4 mM). e, Bulk 13C abundance (atom%) increased (P = 2 × 10−4; ANOVA) over 6 days in living cultures (n = 3; orange circles) incubated with 10% 13C-bicarbonate, ferrihydrite, and daily addition of sulfide (1 mM). f, Bulk 13C abundance increased (P = 0.003; ANOVA) over 6 days in living cultures (n = 3) incubated with 10% 13C-bicarbonate, ferrihydrite, and two spikes of FeS (approximately 1 mM). e,f, Asterisks denote significant differences (P < 0.05; two-sided Student’s t-test) on day 6. Dashed lines indicate natural 13C levels. a–f, Arrows indicate sulfide or FeS addition. g,h, Nano-scale secondary ion mass spectrometry (NanoSIMS) images showing cellular biomass (g; 12C14N–) and 13C content (h) of cells at day 0 and 5. Colour scale refers to 13C atom%. Scale bars correspond to 5 μm. i, Dot plots displaying the 13C content of individual cells analysed in a representative NanoSIMS field view at day 0 (n = 81 cells) and 5 (n = 80 cells). Box plots show the median, the 25th and 75th percentiles of 13C content of cells; whiskers extend to 1.5 times the interquartile range from the first and third quartiles.
After providing physiological evidence for microbial reduction of extracellular iron(iii) being coupled to sulfide oxidation (MISO) in D. alkaliphilus, we revealed the differential activity of candidate genes of this metabolism by comparative transcriptomics under various ferrihydrite-amended and ferrihydrite-free growth conditions. Sulfide oxidation is most probably achieved by the reversal of the canonical dissimilatory sulfate reduction pathway (Fig. 5 and Extended Data Fig. 3), as proposed for D. alkaliphilus growing under nitrate-reducing conditions27 and its close relatives, Electrothrix cable bacteria, which grow by oxygen-dependent sulfide oxidation41. All genes required for this pathway had substantial transcription in D. alkaliphilus during MISO, and many showed significant upregulation (adjusted P value (Padj) < 0.05) compared with cultures grown with formate as electron donor. Additionally, D. alkaliphilus may oxidize sulfide in a two-step process in which sulfide is initially converted into zero-valent sulfur in the periplasm, which is then either disproportionated or transported into the cytoplasm to enter the reverse Dsr pathway27. Multiple genes proposed for this pathway were transcribed and/or upregulated during MISO growth, including those encoding Sqr, the sulfur-transferring membrane protein YeeE, the rhodanese Rhd and the sulfur transferase TusA.
D. alkaliphilus uses a reversed canonical dissimilatory sulfate reduction (Dsr) pathway to oxidize sulfide to sulfate (blue protein labels). Sulfide reacts with DsrC to form DsrC trisulfide (DCT), which is oxidized stepwise to sulfate via DsrAB, AprAB and Sat. The reducing equivalents are transferred to iron(iii) oxides via EET mechanisms (green protein labels) facilitated by MHCs. D. alkaliphilus also has genes encoding type IV pili (grey protein labels). The tight adhesion (Tad) pilus may drive the cell adherence to the surface of solid iron oxides and/or FeS. The transcription level of each gene during MISO is indicated by mean reads per kilobase of transcript per million mapped reads (RPKM) of four replicates corresponding to the colour in the legend. Blue and green dots show significantly increased gene transcription during sulfide/FeS oxidation and ferrihydrite reduction, respectively. The identifier, full name, RPKM, statistical significance of differential transcription and genomic arrangement for each gene are provided in Extended Data Fig. 3 and Supplementary Tables 2–5. Electron flow and proton translocation are indicated by red arrows. Black dotted arrows indicate putative sulfur redox reactions with unclear enzymology. APS, adenosine-5′-phosphosulfate, Qox, oxidized menaquinone; Qred, reduced menaquinone; FoxFeR, incubation of D. alkaliphilus under formate-oxidizing and ferrihydrite-reducing conditions; FeSoxFeR, incubation of D. alkaliphilus under FeS-oxidizing and ferrihydrite-reducing conditions; Sdisp, incubation of D. alkaliphilus under S(0) disproportionation conditions; SoxFeR, incubation of D. alkaliphilus under sulfide-oxidizing and ferrihydrite-reducing conditions; SoxNR, incubation of D. alkaliphilus under sulfide-oxidizing and nitrate-reducing conditions.
After the transfer of sulfide-derived electrons to the membrane quinone pool via the proposed sulfide oxidation pathways, their flow to extracellular ferrihydrite is central to the activities and energy conservation mode of MISO. We propose that this extracellular electron transfer (EET) is facilitated by a network of multi-haem cytochromes (MHCs) at the inner membrane, periplasmic space and outer membrane, similar to mechanisms shown for the model iron(iii)-respiring bacteria Geobacter sulfurreducens and Shewanella oneidensis42,43,44. During MISO growth, D. alkaliphilus showed upregulated transcription of 13 out of 46 predicted MHC genes compared with nitrate-reducing and S(0)-disproportionation growth conditions (Extended Data Fig. 4). The most prominent MHC, upregulated by more than 100-fold (Supplementary Fig. 12), was predicted to be extracellular, and to share structural homology and haem arrangement with OmcS from Geobacter (Extended Data Fig. 5). This OmcS homologue may facilitate EET to iron oxides (Fig. 5, Supplementary Text and Supplementary Fig. 13), as observed in G. sulfurreducens28,45. Another differentially transcribed MHC, also predicted to be extracellular, is homologous to OmcA, which is involved in anaerobic iron reduction in Shewanella46. The remaining MHCs are predicted to be located in the periplasm or associated with the cytoplasmic membrane. One of the latter includes a homologue of a cytochrome bc complex CbcL that is essential for the growth of G. sulfurreducens on iron(iii) oxides47. As proposed for Geobacter, the periplasmic MHCs may facilitate EET by channelling electrons to the extracellular MHCs30; the CbcL homologue may pass electrons from the quinone pool to periplasmic MHCs, while generating a proton motive force via a scalar mechanism47. The involvement of these MHCs in the reduction of iron(iii) oxides was further supported by their increased transcription in D. alkaliphilus grown with formate and ferrihydrite (Extended Data Fig. 4). D. alkaliphilus also upregulated transcription of genes for tight adherence (Tad) type IV pilus assembly proteins during iron(iii) reduction. Transcription of pilA, encoding the main component of another type IV pilus, was high across all growth conditions but not increased during iron(iii) reduction. These proteins were suggested to have a role in EET by forming electrically conductive pili48 or facilitating the surface attachment and/or the secretion of extracellular MHCs49,50,51,52. Under MISO growth conditions, genes involved in Wood Ljungdahl pathways ranked among the top 30% highly transcribed genes (Supplementary Fig. 14), suggesting a role in autotrophic carbon fixation.
Ecological and biogeochemical impacts
Experimental validation of MISO in D. alkaliphilus expands the known physiologies of bacteria, and broadens the diversity of microbial sulfide oxidation pathways beyond what was proved possible in the absence of oxygen. Unlike oxidation of S(0) with dissolved iron(iii) by acidophiles26, MISO relies on solid phase iron(iii) oxides, which represent the most prevalent form of iron in natural habitats with mildly acidic to mildly alkaline pH. Our data suggested that MISO utilizes an elaborate multi-haem apparatus to conduct EET to iron(iii) oxides, similar to iron-reducing model organisms. The proposed sulfide-fuelled EET model provides a mechanistic basis for several underexplored sulfide oxidation observations, including those driven by manganese oxides15, coupled with electricity generation in microbial fuel cells53,54,55, and supporting direct interspecies electron transfer to methanogens56,57. The genomic potential to perform MISO appears as a conserved trait within the Desulfurivibrionaceae family (Extended Data Fig. 6). Desulfurivibrionaceae members with the MISO gene set occupy a wide range of habitats in marine (for example, sediments and hydrothermal vents), freshwater (for example, aquifers and lake sediments), terrestrial (for example, wetlands and soils) and engineered (for example, microbial fuel cells) environments. These observations suggest that MISO-performing Desulfurivibrionaceae members are biogeochemically relevant in diverse ecological contexts. Given the presence of genomic potential for MISO in microbial lineages beyond Desulfurivibrionaceae (Fig. 2, Supplementary Fig. 4 and Supplementary Text), we envisage an even broader environmental distribution of MISO capacity.
This study adds a new dimension to the understanding of the interplay of sulfur and iron biogeochemistry in anoxic environments. Although re-oxidation of sulfide with iron(iii) oxides has previously been regarded to be a strictly chemical reaction, we demonstrate here that MISO can not only outperform chemical oxidation of sulfide at environmentally relevant concentrations, but also acts on solid phase FeS-sulfide, which known to be chemically inert towards iron(iii) oxides38. In this context, FeS, which is widespread in anoxic environments, offers a distinct ecological niche for MISO bacteria, largely devoid of competition from chemical oxidation. Thermodynamic modelling indicates that MISO is exergonic over a range of ecologically relevant pH values and substrate conditions. This contrasts with canonical reduction processes of iron(iii) oxides, which become thermodynamically unfavourable at alkaline pH19. MISO is likely to overcome such energy limitation by exploiting highly reactive sulfide that maintains its reducing power over a wide pH range. The free energy yielded from MISO is sufficient to support microbial growth and power sulfate formation at a rate resembling that of sulfate reduction in environments such as marine sediments (Supplementary Text). These unique features suggest that MISO bacteria may directly facilitate environmental sulfate formation from sulfide or FeS with iron(iii) oxides, bypassing the need for S(0) disproportionation. Including MISO in the existing biogeochemical model could help explain the observed widespread geochemical pattern of sulfide removal and sulfate formation in anoxic, iron-rich marine, freshwater and terrestrial environments14,20,22,58,59,60 (Supplementary Text). A notable example of such phenomena is the occurrence of high oxidation rates of sulfide, produced from dissimilatory sulfate reduction, back to sulfate in anoxic layers of marine sediments, where typical oxidants for sulfide oxidation, such as oxygen and nitrate, have been depleted, leaving iron(iii) oxides as the primary oxidants20. On the basis of a first estimate, MISO could account for 1–7% of total sulfide oxidation to sulfate in marine sediments on a global scale, owing to the large flux of reactive iron from terrestrial runoff to the sea (Supplementary Text). The MISO metabolism may therefore be globally important in modulating the oxidation and reduction of sulfur, with broader implications for Earth’s biogeochemical cycles and climate.
Methods
Phylogenetic framework and HMMs of sulfur-cycling proteins
To discern sulfur-cycling genes or proteins from their functionally divergent homologues, phylogenetic analysis was conducted for 116 sulfur-cycling proteins (Supplementary Table 1). For each protein family, sequences of enzymes with biochemically validated functions, including related sequences of enzymes with divergent functions (outgroups), were identified by literature surveys and recovered from SwissProt61. Additional homologues of experimentally validated proteins in KEGG prokaryotic genomes were retrieved using KEGG BLAST Search (https://www.genome.jp/tools/blast/; E value: 10−4). Distant homologues that did not align properly with biochemically characterized proteins (alignment length covered <50% of query and target length) were removed. The resulting homologues were de-replicated using CD-HIT v4.8.1 (ref. 62), with longest sequences retained as representatives. For computational efficiency, different clustering identity thresholds (75–95%) were chosen for de-replication to ensure the total number of representative sequences for phylogenetic analysis did not exceed 500. The genome context of the representative KEGG homologues was analysed by retrieving genes located in a distance of fewer than n genes (n = 7–15), followed by annotation using biochemically characterized gene clusters based on BLAST analysis63. All representative KEGG homologues were further aligned with biochemically validated proteins and outgroups using Muscle v3.8.1551 (ref. 64). Poorly aligned regions were excised using TrimAl v1.4.rev15 (ref. 65). Protein phylogeny was inferred from the trimmed alignment using FastTree v2.1.7 (ref. 66) with -wag and -gamma options. Statistical support for each branch of the tree was estimated by nonparametric bootstrap (n = 100).
Information on reference sequences from biochemically verified proteins (for example, ingroup/outgroup, conserved residues or motif) and genomic contexts of all homologues were mapped on the tree. To identify monophyletic, orthologous clades within each tree, interior nodes of the annotated tree were scrutinized using the following criteria: (1) bootstrap support over 70%; (2) presence of at least one biochemically verified ingroup protein and absence of outgroup proteins; and (3) consistent gene neighbouring patterns and biochemical traits (thatis, catalytic residues and PFAM domain composition) among its members. All descendants of the identified clade were regarded as functional orthologs of the biochemically verified protein. If possible, existing definitions of orthologous clades from previous phylogenetic analysis of sulfur-cycling proteins was preserved, including the well-recognized clades in the phylogeny of DsrAB10 and Sqr67. For proteins for which the biochemically validated ingroup proteins formed polyphyletic groups, multiple monophyletic clades were proposed to fulfill our criteria.
To leverage our phylogenetic framework for large-scale homology searches, sequences from the defined monophyletic clades of sulfur-cycling proteins were used to build HMMs. A cut-off that optimizes the sensitivity and specificity of homology search was calculated for each HMM using receiver operating curve (ROC)68. This cut-off was embedded in the HMMER profile HMM file as the gathering threshold of the model (HMMER User’s Guide, p. 108; ref. 69). The performance of the newly developed HMMs was compared with that of six published sets of HMMs for sulfur metabolism genes, including those from KoFam70, TIGRFAM71, PFAM72, metabolicHMM73, DiSCo74, Teng et al.75 and HMS-S-S76. This was accomplished by querying each HMM against the phylogeny-curated protein dataset using hmmsearch in HMMER v3.2.1 with a predefined cut-off (http://hmmer.org/). The performance of the various HMM sets in detecting sulfur-cycling genes and proteins was assessed in terms of specificity, sensitivity, and F score (Supplementary Text). F score balancing both precision and recall of the homology detection was calculated using F score = 2 × (precision × recall) / (precision + recall).
Sulfur-cycling genes in bacterial and archaeal genomes
To provide a comprehensive overview of sulfur metabolism across bacteria and archaea, the phylogeny-derived HMMs were searched against all genomes in GTDB release 95 (ref. 23) using hmmsearch with the --cut_ga option. Each retrieved homologue was then searched against the full set of phylogeny-derived HMMs using hmmscan with --cut_ga, and annotated as the HMM showing the highest score. For initial screening, a subset of genes (n = 42) was selected as markers for specific sulfur metabolisms if the gene: (1) has been widely recognized as a marker for a specific sulfur metabolism, (2) encodes a catalytic subunit essential for the activity of enzymatic complex; or (3) on its own confers a specific sulfur redox transformation (see justification for each of selected genes in Supplementary Table 1). The retrieved homologues were further curated using our reference phylogeny of sulfur proteins. Specifically, the GTDB homologues were aligned with sequences contained in our reference phylogeny using Muscle. A maximum-likelihood tree was reconstructed from the alignment trimmed by TrimAl. The tree was overlaid with biochemical information and data on the genomic context of sulfur genes, and visualized using ggtree77. The physiological role of the GTDB homologues was interpreted on the basis of their evolutionary relationship with biochemically validated proteins and genome context. To predict the dissimilatory iron(iii) reduction potential, GTDB genomes were screened for marker genes involved in EET on the basis of homology search and/or motif analysis. Homologues of iron(iii) reduction genes with established HMMs in FeGenie database (https://github.com/Arkadiy-Garber/FeGenie/tree/master/hmms/iron/iron_reduction) were retrieved using hmmsearch from HMMER v3.2.1, with cut-off recommended by FeGenie (https://github.com/Arkadiy-Garber/FeGenie/blob/master/hmms/iron/HMM-bitcutoffs.txt). Additionally, homologues of MmcA gene, which is involved in dissimilatory iron(iii) reduction in Methanosarcina acetivorans78, were extracted using BLASTP on the basis of an e-value of 10−4. The outer membrane MHCs responsible for EET with metal oxides in anaerobic methanotrophs79 and putative electroactive bacteria80 were recognized on the basis of the following: (1) the presence of four or more haem-binding motifs (CXXCH); and (2) their predicted outer membrane or extracellular localization, as determined by DeepProLoc v1.0 (ref. 81).
Annotation and metabolic reconstruction of specific sulfur-cycling microbial lineages
The genomes of microbial lineages of interest were downloaded from the GTDB database. The protein-coding genes were predicted from the genome using Prodigal v2.6.3 with default setting. The predicted genes were annotated using KoFam70, PFAM72, and the EggNOG82 database. Additional metabolic pathways were predicted using HMMs (Supplementary Table 6) downloaded from dbCAN83, metabolicHMM73, CANT-HYD84, MicRhoDE85 and FeGenie86. For HMM-based annotation, the HMMs were used as queries to search against microbial genomes using hmmsearch from HMMER v3.2.1, with cut-off recommended by each database (-T, -domT or -cut_ga options). The cellular localization of the protein was predicted using Signalp v6.0 (ref. 87). The completeness of the KEGG metabolic pathway was calculated on the basis of the definition of each module. The KEGG module is defined with a logic expression of K numbers that records the composition of enzymes in the pathway. A particular metabolic module was considered to be present in the genome when: (1) the diagnostic/marker genes of the module were detected; and (2) the overall completeness of the pathway module was >70%. The environmental distribution of the GTDB species was retrieved by searching their GTDB species name in the Sandpiper database88. The occurrence of the GTDB species across biomes was downloaded as CSV from Sandpiper (https://sandpiper.qut.edu.au/) and further visualized with R v4.1.0.
Thermodynamic modelling
The Gibbs free energy associated with iron(iii)-dependent sulfur oxidation at environmentally relevant conditions was estimated by following the guidelines described previously89. In brief, the actual Gibbs free energy of reaction (ΔGr) was calculated using:
where ΔGr0 refers to the standard Gibbs free energy of reaction, given in kJ mol−1; R and T are the universal gas constant (8.314 J K−1mol−1) and the temperature in Kelvin, respectively; and Qr is the reaction quotient. ΔGr0 values were calculated from the values of the standard Gibbs free energy of formation (ΔGf0) of reactants and products (Supplementary Table 7). Values of Qr were determined from the activity (ai) and the stoichiometric coefficient (vi) of the ith chemical species involved in the reaction using:
The activity of the solvent (that is, pure water) and the solids (that is, ferrihydrite and FeS) were taken to be 1. The activity of dissolved ions was related to the concentration (Ci) using:
where γi denotes the activity coefficient; Ci0 represents the standard state concentration (usually 1 M). γi for cations (that is, Fe2+) and anions (that is, HS−, S2O32− and SO42−) in solutions of different ionic strength were retrieved from Amend et al.89. Sulfide speciation in aqueous phase across a range of pH was determined from the pH, and pKa1 (7.04) and pKa2 (11.96) of hydrogen sulfide.
Synthesis of ferrihydrite and poorly crystalline FeS
Synthetic ferrihydrite was prepared by titrating 1 M NaOH (Sigma Aldrich) into 0.1 M aqueous solution of FeCl3 6H2O (Carl Roth) under vigorous stirring until pH 7.5 was reached, as described90. The suspension was centrifuged (Centrifuge 5804 R, Eppendorf) at 4 °C, 12,857g and the ferrihydrite nanoparticles were washed thoroughly with deionized water to remove traces of chloride. The pellets were then freeze-dried (Alpha 1-4 LSCbasic, Christ) and stored at −20 °C for no longer than 3 weeks before use. The mineralogy was determined by LabRAM HR800 Raman microscope (Horiba Jobin-Yvon) equipped with a 532-nm neodymium-yttrium aluminium garnet laser and either 300 or 600 grooves/mm diffraction grating. Iron monosulfide (FeS; 30 mM) was prepared by mixing equal volume of 60 mM aqueous solution of Na2S.9H2O (Acros Organics) with 60 mM aqueous solution of FeCl2.4H2O (Sigma Aldrich) in an anaerobic chamber (Coy Lab) with 95% N2 and 5% H2 (O2 < 1 ppm) atmosphere. The initially precipitated FeS is often designated as ‘amorphous FeS’ or ‘poorly crystalline FeS’91. The dissolved sulfide in the FeS stock is less than 50 µM. The FeS solution was freshly prepared and used on the same day.
Cultivation of D. alkaliphilus DSM 19089
D. alkaliphilus (DSM 19089, ATH2) was purchased from the German Collection of Microorganisms and Cell Cultures GmbH (DSMZ). The bacterium was cultivated at room temperature in an alkaline mineral medium (pH 9.3) containing 3 g NaCl (Carl Roth), 0.25 g K2HPO4 (Merck), 6.5 g Na2CO3 (Carl Roth), and 15 g NaHCO3 (Sigma Aldrich) per liter of medium. After autoclaving, the medium was cooled down under N2 atmosphere and supplemented aseptically with 1 ml liter−1 of following components (all stored under anoxic conditions): 4 M NH4Cl (Sigma Aldrich), 1 M MgCl2 (Sigma Aldrich), trace element solution, Se-W solution, and four different vitamin solutions (DSMZ medium 1104). The culture was routinely grown under nitrate-reducing, sulfide-oxidizing conditions in 500 ml Schott bottles27, with 2 mM Na2S 9H2O and 1.2 mM KNO3 (Sigma Aldrich). This yielded a culture with an optical density at 600 nm (OD600) of ~0.040, corresponding to a cell density of ~1.3 × 108 cells per ml. To test alternative growth modes, five incubation experiments were conducted, each supplemented with different electron donors and acceptors (details provided below). For all experiments, regularly maintained cultures (30 ml) that have been depleted in sulfide (< 100 µM) and nitrate (< 10 µM) were used as inoculum. Incubations were set up in 60 ml serum bottles and sealed with butyl rubber stoppers in the anaerobic chamber (N2:H2 = 95:5). Each culture was then flushed with pure N2 to remove H2 in the headspace, and incubated in the dark at room temperature. All incubations, abiotic and biotic controls from each experiment were set up in triplicates.
-
(1)
Incubations with sulfide and nitrate. The incubations were set up by supplying 2 mM sulfide and 2 mM nitrate to 30 ml pre-growns cells in 60 ml serum bottles. Sulfide and nitrate was spiked using syringes flushed with pure N2. The growth was monitored by phase-contrast microscopy and by the measurement of sulfide and sulfate over 3 days.
-
(2)
Incubations with elemental sulfur. The incubations were initiated by adding 0.1 g elemental sulfur in 3 ml MilliQ water (Sigma Aldrich) to each of the serum bottles, followed by autoclaving at 110 °C for 60 min. After sterilization, 30 ml pre-grown cells were inoculated into the S(0) suspension (approximately 94 mM) and incubated under an N2 atmosphere for 15 days. Microbial activity was monitored by measuring sulfide and/or sulfate.
-
(3)
Incubations with ferrihydrite and formate. Synthetic ferrihydrite (0.2 g) was ground into fine particles with an agate mortar and pestle before being added to the culture. Assuming ferrihydrite has a composition92 Fe(OH)3, the final concentration of Fe(iii) was approximately 62 mM. Formate was spiked anoxically using a syringe to a final concentration of 10 mM. To test the coupling of ferrihydrite reduction and formate oxidation, parallel cultures were set up with either ferrihydrite or formate alone. The culture activity was monitored by measuring total Fe(ii) and formate concentrations over a 15 day incubation.
-
(4)
Incubations with ferrihydrite and poorly crystalline FeS. Synthetic ferrihydrite (0.2 g) was supplied to the cultures as described in the incubation (3). To amend poorly crystalline FeS, 1 ml stock solution of freshly prepared FeS (30 mM) was anoxically spiked to the cultures using syringes, resulting in a final FeS concentration of 1 mM. Abiotic controls were prepared using 30 ml autoclaved cells as inoculum to test for chemical reactions between ferrihydrite and poorly crystalline FeS. Biotic controls amended with either ferrihydrite or FeS were set up to assess the impacts of residual sulfide and/or nitrate on culture activity. Cultures were sampled daily over 5 days for sulfate and total Fe(ii) measurements.
-
(5)
Incubations with ferrihydrite and dissolved sulfide. The cultures were prepared similarly as incubation (4), but with dissolved sulfide replacing FeS. Due to the rapid chemical reaction between dissolved sulfide and ferrihydrite, dissolved sulfide was anoxically spiked daily at a concentration of 1 mM using N2-flushed syringes. Abiotic controls and the sulfide-only biotic controls received dissolved sulfide at the same concentration and frequency. To trace the transformation of S and Fe over 5 days, subsamples were taken daily for measurement of S(0), total Fe(ii), sulfate, and Cline-extractable sulfide before the addition of sulfide. The consumption of dissolved sulfide in the cultures was monitored by sampling at 2, 10, 20, 35, 50 and 70 min after the spike of sulfide. The kinetics of sulfide consumption were modelled as a first-order reaction. The rate constant was estimated using the exponential decay model in the drm() function from the drc R package93. To compare sulfate formation patterns with and without sulfide, cultures incubated with ferrihydrite and sulfide were sampled for sulfate measurement following two phases after the 1st sulfide spike. During phase I, detectable sulfide was present in the culture, and the samples were collected at 0, 11, 21, 37, 53 and 70 min of the incubation. Phase II, spanning the next 23 h, began once sulfide was depleted, with samples taken at 3, 5.5, 8.33, 12.33, 20.25 and 24 h. As a control for phase II, cells were incubated with chemically sulfidized ferrihydrite. Specifically, 1 mM sulfide was firstly added to 0.2 g ferrihydrite (approximately 62 mM) with 30 ml autoclaved cultures for chemical reaction. After 70 min, the reaction mixture was centrifuged (12,857g; room temperature) under anoxic conditions, and 30 ml of active cells were inoculated to resuspend the solid phase compounds (for example, FeS and S(0)) produced by chemical reaction between sulfide and ferrihydrite. Samples were collected from cultures for sulfate measurement at the same time intervals as those in phases I and II.
To test whether the microbial process can outperform the chemical process in transforming sulfide with ferrihydrite, the incubation (5) was repeated using ca. 50 µM sulfide instead of 1 mM. In this experiment, a small amount of sulfide was spiked three times at 1.5-h intervals into ferrihydrite-amended cultures, abiotic controls, and sulfide-only biotic controls. After each spike, subcultures (~ 0.3 ml) were collected at 2, 5, 10, 15, 20, and 25 min for dissolved sulfide measurements. Two biological replicates were performed for each treatment. To verify the reproducibility of the observed sulfide consumption pattern, incubations were conducted using inocula at different cell densities (OD600 of 0.042, 0.075, and 0.086). To quantify the transformation of spiked sulfide during the incubation, independent cultures were set up using an inoculum with an OD600 of 0.072 and supplied with eight spikes of sulfide. Ferrihydrite-amended cultures, abiotic controls, and sulfide-only controls received ca. 50 µM sulfide at 1.5-h intervals, whereas ferrihydrite-only biotic controls were spiked with anoxic water. Three replicate incubations were performed for each treatment. Subsamples were taken every three hours for concentration measurement of S(0), total Fe(ii), sulfate and Cline-extractable sulfide.
Chemical analysis of metabolites
To monitor the dynamics of metabolites in the incubation experiments, subsamples of the culture were taken periodically with sterile syringes flushed with pure N2 as described above. HCl-extractable Fe(ii) was determined by adding 0.1 ml sample aliquots to 0.2 ml 0.75 N HCl. The sample was immediately centrifuged for 15 min at 12,044g. Fe(ii) in the resulting 0.5 N HCl was measured using the ferrozine assay. Previous studies have shown the 0.5 N HCl treatment allowed quantitative extraction of the solid phase Fe(ii) associated with the surface of iron oxides, Fe(ii) from FeS, and the dissolved Fe(ii) in the Fe/S system7,94. Therefore, we referred to HCl-extractable Fe(ii) as total Fe(ii).
Aqueous and total sulfide were determined using spectrophotometric methods. To measure dissolved sulfide, approximately 0.3 ml subculture was filtered through a 0.2 µm membrane (CHROMAFIL). The dissolved sulfide in the filtrate (0.1 ml) was fixed by 0.25 ml 3% w/v zinc acetate dihydrate (Sigma Aldrich), followed by quantification using the Cline method95. The filtered sample from the incubation with ferrihydrite and 1 mM dissolved sulfide showed black colour, indicating the formation of FeS particles smaller than 0.2 µm. The sulfide associated with this FeS fractionation was approximated as HCl-extractable Fe(ii), assuming a 1:1 stoichiometry. The total sulfide was determined as Cline-extractable sulfide. The Cline reagent contains 6 N HCl that dissolves some solid sulfides (for example, freshly formed FeS), and thus the Cline-extractable sulfide comprises dissolved sulfide and HCl-reactive solid phase sulfide. Total sulfide in the Fe/S system is typically determined as acid volatile sulfide. Acid volatile sulfide was not analysed here owing to the large uncertainties inherent to this methodology91,96.
Sulfate and formate concentrations in the incubations were determined by capillary electrophoresis techniques. Sample preparation for sulfate measurement involved fixation of 100 µl subsample with 10 µl 3% w/v zinc acetate, dilution with 890 ul MilliQ water, filtration through a 0.2 µm membrane, and addition of 1 mM chlorate as the internal standard. The standards were prepared by adding defined amounts of sulfate (Sigma Aldrich) to the alkaline medium, followed by the same treatment procedure as described for samples. The sulfate content in the prepared samples/standards was measured using an Agilent 7100 capillary electrophoresis system (Agilent Technologies), equipped with a capillary (72 cm × 72 µm internal diameter; Agilent Technologies) and a diode array UV-vis detector (DAD). Electrolytes for anion separation contains 2.25 mM pyromellitic acid (Sigma Aldrich), 1.6 mM triethanolamine (Sigma Aldrich), 0.75 mM hexamethonium hydroxide (Sigma Aldrich), and 6.5 mM NaOH97 at pH 7.8 ± 0.1. Anion separation was implemented at a voltage of −30 kV. The data were acquired through indirect UV detection at a wavelength of 350 nm with a bandwidth of 60 nm, and a reference wavelength of 245 nm with a bandwidth of 10 nm. For the formate measurement, 900 µl MilliQ water was added to 100 µl samples/standards (Sigma Aldrich), which were then filtered through a 0.2 µm membrane. l-malate (Sigma Aldrich) was added to the filtrate as the internal standard. Organic Acids Buffer for capillary electrophoresis (pH 5.6; Agilent Technologies) was used as electrolytes, and the separation conditions, including DAD and capillary electrophoresis settings, were configured according to manufacturer’s instructions. All electropherogram data were analysed with the Agilent ChemStation.
Elemental sulfur was measured using high performance liquid chromatography (HPLC). One-hundred microlitres of sample was fixed with 10 µl of 3% w/v zinc acetate. Then, 300 µl chloroform was added, and the mixture was shaken at 500 rpm for 1 h. The elemental sulfur in chloroform phase was then measured using a Dionex UltiMate 3000 UPLC system, equipped with an UltiMate 3000 pump (0.2 ml min−1), a column Compartment (25 °C), a column Waters ACCQ-TAG ULTRA C18 1.7 µm × 2.1 × 100 mm, and an UltiMate 3000 Variable Wavelength Detector (UV) (wavelength 254 nm). The isocratic elution with 100% methanol was applied. With these adjustments, the peak appeared after 3.4 min. Data were analysed with Dionex Chromeleon software.
Microscopy of D. alkaliphilus incubated with ferrihydrite and sulfide
For scanning electron microscopy (SEM), transmission electron microscopy (TEM) and fluorescence microscopy, cultures incubated with ferrihydrite and sulfide (daily spike of 1 mM) for 5 days were fixed in 2% glutaraldehyde or 2.3% formaldehyde, respectively. For SEM imaging, solid iron phase iron was allowed to settle without centrifugation, carefully washed with MilliQ water, and transferred to 100% ethanol. Samples were then dried using rapid chemical drying with hexamethyldisilazane and mounted on aluminium stubs with double-sided sticky carbon tape and sputtered with Gold (JEOL JFC-2300HR). The images were taken with a Scanning Electron Microscope (JEOL IT 300 LAB6EOL) with Secondary Electron Detector (SED) and Backscattered Electron Detector (BED-C) at 20 kV.
For TEM imaging, cultures were treated with a solution containing 50 g l−1 sodium dithionite, 0.2 M sodium citrate and 0.35 M acetic acid (hereafter termed dithionite solution) as previously described98. After dissolution of solid iron phase, cells were pelleted at low speed (2,300g) to minimize shear forces and washed with MilliQ water before suspending cells in MilliQ water. For negative staining, 4 µl of sample was incubated for 1 min on a formvar-filmed and carbon-coated grid (200 mesh, Cu) and excess liquid was removed with a filter paper. A drop of stain (2.5% gadolinium acetate) was applied and immediately removed. Samples were examined in a TEM EM 900 N (Zeiss) at 80 kV.
For fluorescence microscopy, the formaldehyde-fixed cultures were resuspended and a subsample was filtered onto a 0.2 µm pore size polycarbonate membrane (Millipore). Cells on the filter were stained with a 1× SYBR Green solution, and images were acquired using a epifluorescence microscope (Zeiss Axio Imager M1 with an AxioCam MRm).
Monitoring the growth of D. alkaliphilus during incubation experiments
Growth was monitored by cell counting for cultures incubated under 4 different conditions: (1) ferrihydrite (approximately 62 mM Fe equivalent) and periodic spike of approximately 50 µM dissolved sulfide (sulfide was spiked 40 times over 5 days, with one spike every hour and 8 times per day); (2) ferrihydrite (approximately 62 mM Fe equivalent) and daily spike of 1 mM sulfide; (3) ferrihydrite (approximately 62 mM Fe equivalent) and periodic spike of FeS (approximately 1 mM Fe equivalent); and (4) nitrate (4 mM) and 2 spikes of sulfide at concentration of 1–2 mM. The setup of the cultures and controls was the same as described in ‘Cultivation of alkaliphilus DSM 19089’ except that a lower starting cell density (3–5 × 107 cells per ml) was used. During each of the incubation experiments, subcultures (450 µl) were sampled periodically and preserved in 2.3% formaldehyde (final concentration). Before counting, 500 µl dithionite solution was added to 50–100 µl of fixed cells to dissolve the FeS and ferrihydrite particles. After dissolution of solid iron phase (within 10–15 min), 100 µl of each sample was diluted in 900 µl of 1× phosphate-buffered saline (PBS). The suspension was then sonicated using a SONOPULS ultrasonic homogenizer (Bandelin, Berlin, Germany) at 25% power with a cycle setting of 2 for a total of 30 s. Cells were subsequently stained with SYBR Green 1× (ThermoFisher) and incubated for 10 min at room temperature in the dark. Flow cytometric analysis was performed using a CytoFLEX S flow cytometer (Beckman Coulter) equipped with a blue 488 nm laser. SYBR Green fluorescence was detected using a 525/40 nm bandpass filter. A fluorescence threshold was applied on the SYBR Green signal to exclude background events. For each sample, 80–100 µl was measured. Data were gated on SYBR Green–positive cells displaying fluorescence shifts relative to unstained controls to identify the target population (Supplementary Fig. 15). Data were acquired and analysed with the CytExpert 2.6 software (Beckman Coulter). The specific growth rate (k; day−1) was estimated via linear regression analysis of ln(Cellt/Cell0) versus time (day) over an apparent exponential growth phase. Here, Cellt is the cell concentration (in cells per ml) at sampling time t (day).
13C-bicarbonate labelling experiments and isotope analysis
To probe for autotrophic carbon fixation during MISO growth conditions, 13C-labelled bicarbonate (98 atom% 13C; Sigma Aldrich) was added to ferrihydrite-incubated cultures receiving dissolved sulfide (1 mM) or FeS (ca. 1 mM S equivalent), to reach a 10 atom% 13C in the inorganic carbon pool. The dissolved sulfide or solid phase FeS were spiked in the same frequency as for the growth experiment. Abiotic controls for each culture were set up using autoclaved inoculum. To detect 13C content in bulk biomass and in single cells, subcultures were sampled, fixed by formaldehyde (2.3% final concentration), and analysed using elemental analyser-isotope ratio mass spectrometry (EA-IRMS) and NanoSIMS. For EA-IRMS, 1.5 ml of fixed samples that included ferrihydrite and cells were pelleted by centrifugation and washed with MilliQ water, followed by overnight treatment by 0.1 M HCl to remove residual carbonates. The dried cells attached to ferrihydrite particles were weighed (4–6 mg) and transferred to tin cups. Bulk cell carbon isotope ratios (13C:12C) were measured by EA-IRMS (Delta V Advantage) coupled by a ConFlo IV interface to an elemental analyser (EA-Isolink, all Thermo Finnigan). Sample 13C contents were calculated as atomic percentage of 13C in total carbon, following 13C atom% = 13C/(13C + 12C) × 100%. The analytical precision of replicate analyses of isotopically homogeneous international standards was ±0.0001% for 13C atom% measurements.
For NanoSIMS analysis, 0.1 ml formaldehyde-fixed samples that included ferrihydrite and cells were mixed with dithionite solution as described above and incubated for 2 h. After complete dissolution of ferrihydrite, 400 µl of the suspension was transferred onto gold-coated polycarbonate filters (GTTP type, 0.2 µm pore size, Millipore). The filters were gold-coated by physical vapour deposition, utilizing an Agar B7340 sputter coater (Agar Scientific) equipped with an Agar B7348 film thickness monitor (Agar Scientific) for precise adjustment of the coating thickness (150 nm). The filters were incubated for 2 h in 0.1 M HCl to remove residual carbonates and then washed twice in MilliQ water and then air-dried. Filter sections were attached to antimony-doped silicon wafers (7.1 ×7.1 mm, Active Business Company) with a commercially available glue (SuperGlue Loctide).
NanoSIMS measurements were carried out on a NanoSIMS 50 L instrument (Cameca) at the Large-Instrument Facility for Environmental and Isotope Mass Spectrometry at the University of Vienna. Prior to data acquisition, analysis areas were pre-conditioned in situ by rastering a high-intensity, slightly defocused Cs+ ion beam for removal of surface adsorbates and establishment of the steady state secondary ion signal intensity regime with minimum sample erosion. For this purpose, the following sequence of high and extreme low Cs+ ion impact energy (EXLIE) was applied: high energy (16 keV) at 100 pA beam current to a fluence of 5 × 1014 ions cm−2; EXLIE (50 eV) at 400 pA beam current to a fluence of 5 × 1016 ions cm−2; high energy to an additional fluence of 2.5 × 1014 ions cm−2. Data were acquired as multilayer image stacks by scanning of a finely focused Cs+ primary ion beam with 2 pA beam current at approximately 80 nm physical resolution (probe size) over areas between 60 × 60 and 62 × 62 µm2 with 512 × 512 pixel and 1,024 × 1,024 pixel image resolution and a per-pixel dwell time of 5 ms and 1.5 ms, respectively. The detectors of the multicollection assembly were positioned for parallel detection of 12C2−, 12C13C−, 12C14N−, 31P− and 32S-secondary ions. Secondary electrons were detected simultaneously for gaining information about the sample morphology and topography. The mass spectrometer was tuned to achieve a mass resolving power ((MRP) = M/ΔM) of >10,000 for detection of C2− secondary ions.
Measurement data were processed using the WinImage software package provided by Cameca (WinImage V4.8) and the OpenMIMS plugin in the image processing package ImageJ (V1.54p). Prior to data evaluation, images were corrected for detector dead-time and positional variations emerging from primary ion beam and/or sample stage drift. Carbon isotope composition images displaying the 13C/(12C + 13C) isotope fraction, given in atom percent (atom%), were inferred from the C2− secondary ion signal intensity distribution images via per-pixel calculation of 13C12C−/(2 × 12C2− + 12C13C−) intensity ratios. For numerical data evaluation, regions of interest, referring to individual cells, were manually defined on the basis of the 12C14N− and 31P− secondary ion maps as indicators of biomass and verified by the topographical/morphological appearance in the secondary electron images. Biomass aggregates, in which an unambiguous identification of single cells was not feasible, were rejected.
Cells were assessed as being significantly enriched in 13C after incubation in the presence of 13C-bicarbonate if (1) the 13C isotope fraction value was higher than the mean plus 3 standard deviations (σ) of the values determined on the cells from the control (on day 0) and (2) the statistical counting error (3σ, Poisson) was smaller than the difference between the considered 13C enriched cell and the mean value measured on the cells from the control. The Poisson error was calculated from the secondary ion signal intensities (given in counts per region of interest) via
On the basis of these two criteria, all individual cells measured in the 13C incubated sample showed a significant enrichment in 13C.
RNA-seq and transcriptomics
D. alkaliphilus cultures grown under five incubation conditions (as described in ‘Cultivation of D. alkaliphilus DSM 19089’), each in four replicates, were used for comparative transcriptomic analysis. Cultures (30 ml) showing metabolic activity (for example, Fe(ii) production, sulfide consumption or production) were collected in the middle to late stage of incubation experiments. Cells were collected by centrifuging (12,857g; room temperature) under anoxic condition using oak ridge tubes (Thermo Fisher Nalgene) with replacement O-rings for sealing cap (Thermo Fisher Nalgene). The cell pellets were resuspended with 1.5 ml supernatant, and distributed to three lysis matrix E tubes (MP Biomedicals), each with approximately 0.5 ml. The collected cells were immediately frozen with liquid N2, and stored at −80 °C before subsequent analysis. The total nucleic acids were extracted following a phenol-chloroform protocol as described previously99,100. In brief, the sample was lysed for 30 s at a speed of 5.5 m s−1, after mixing with hexadecyltrimethylammonium bromide extraction buffer and phenol-chloroform-isoamyl alcohol (25:24:1) (pH 8.0). The aqueous phase was extracted by centrifugation, and the phenol within was removed by mixing with chloroform-isoamyl alcohol (24:1). The total nucleic acids in the aqueous phase were then precipitated with polyethylene glycol 6000, followed by centrifugation. The pelleted nucleic acids were washed with ice-cold ethanol and dried before resuspension in diethyl pyrocarbonate-treated water. DNA from the total nucleic acids were removed using the TURBO DNA-free kit (Thermo Fisher Scientific).
RNA-sequencing was performed at the Joint Microbiome Facility of the Medical University of Vienna and the University of Vienna (JMF) under project IDs JMF-2311-14 and JMF-2405-05. Sequencing libraries were prepared from rRNA depleted (Ribo-Zero Plus rRNA Depletion Kit, Illumina) RNA samples (NEBNext Ultra II Directional RNA Library Prep Kit for Illumina, New England Biolabs) and sequenced in 2× 100 bp paired-end mode (NextSeq 6000, Illumina), yielding 74.2–303.7 million raw reads per sample. Individual read libraries were quality checked using fastQC v0.12.1 (http://www.bioinformatics.babraham.ac.uk/projects/fastqc/) and quality statistics were merged using multiQC v1.21 (ref. 101). Adapters were trimmed and phiX contamination was removed using BBDuk (part of BBMap v39.06). Reads were k-trimmed from the right with a kmer of 21, minimum kmer of 11 and hamming distance of two along with the tpe and tbo options. Quality trimming was performed from the right with a q-score of 28 to a minimum of 50 bases in length (https://sourceforge.net/projects/bbmap/). The quality filtered reads were aligned to the reference genome of D. alkaliphilus (NC_014216.1) using BBMap with a mapping identity of 99% and with ambiguous reads assigned to the best location (that is, counted only once for duplicated genes). FeatureCounts (part of SubRead 2.0.6 (ref. 102)) with reverse-stranded and –countReadPairs were used to generate counts tables with the resulting alignments based on gene call locations by prodigal v2.6.3 (ref. 103). Counts tables were analysed using DESeq2 release 3.19 (ref. 104) to calculate the RPKM and to determine statistical significance of differential transcription between treatment groups. All P values are adjusted for multiple comparisons using the Benjamini–Hochberg method105.
Quantitative PCR with reverse transcription (RT–qPCR) was performed to verify the upregulated transcription for the MHC gene DA_402 under iron-reducing conditions. Primers DA_402_998F (5′-TTCCCAATCGGGGCGAATAC-3′) and DA_402_1081R (5′-TGGCCTCGGTATAGAGGGTC-3′) were used to target DA_402. Primers recA_79F (5′-TTCGGCAAAGGCTCCATCAT-3′) and recA_221R (5′-TCCGGCCCATATACCTCGAT-3′) were used to quantify the transcription level of the house-keeping gene recA (DA_1926) encoding the DNA recombination protein. Primers for both genes were newly designed using Primer-Blast106. For RT–qPCR, DNA-free RNA was first reverse transcribed to cDNA using SuperScript III reverse transcriptase according to the manufacturer’s instructions. The absolute abundance of transcripts from DA_402 and recA were quantified by quantitative PCR using cDNA as a template. Purified PCR products of gene DA_402 and recA amplified from genomic DNA of D. alkaliphilus were used as quantitative PCR standards. The PCR reactions were prepared in triplicates and run at 95 °C for 3 min, followed by 40 cycles of 95 °C for 30 s, 60 °C for 30 s, and 72 °C for 45 s, on the Thermal Cycler with CFX96 Real-Time System (Bio-Rad). The RT–qPCR data were acquired and analysed using CFX Maestro software (Bio-Rad). The transcription level of DA_402 was compared between treatments after normalization with that of recA. The statistical significance of differential transcription between treatments were determined via Student’s t-test.
Structure prediction and phylogenetic analysis of multi-haem c-type cytochromes in D. alkaliphilus
D. alkaliphilus proteins with more than one haem-binding motifs (CXnCH; n = 2 to 5) were considered MHCs107. The haem-binding motifs in protein sequences were counted using regex expressions in the Python re package. The subcellular localization of all putative MHCs (n = 46) from D. alkaliphilus was predicted using PSORTb v3.0 (ref. 108) and DeepLocPro v1.0 (ref. 81). Prediction from DeepLocPro was used for the proteins for which PSORTb returned ‘Unknown’. The transcription levels of MHCs were compared between different incubation experiments on the basis of RPKM values. The statistical significance of differential transcription was assessed as described in the ‘RNA-Seq and transcriptomics’ chapter. The most highly transcribed extracellular MHC (DA_402) during MISO was further selected for structure prediction and phylogenetic analysis. The structure of the DA_402 monomer and oligomer were predicted using AlphaFold2 v2.3.2 at the COSMIC2 science gateway. The leading signal peptide, predicted using SignalP 5.0 (ref. 109), was cleaved from the protein sequence before structure modelling. For comparison, the cryo-EM structure of OmcS from G. sulfurreducens was retrieved from the Protein Data Bank (PDB) database (6EF8). The protein sequence of DA_402 was aligned to OmcS using the T_coffee alignment tool110. The structure-structure similarity between DA_402 and OmcS was calculated using an online TM-align tool and DaliLite.v5 (ref. 111,112,113). The haem-binding sites in DA_402 and OmcS were visualized using MacPyMOL v.1.7.4 (https://pymol.org). The haem was docked to the target haem-binding site in DA_402 using AutoDockTools v1.5.7 (ref. 114) and AutoDock Vina 1.1.2 (ref. 115). To conduct phylogenetic analysis of DA_402, homologues of DA_402 were retrieved from the KEGG database using Blastp with an E value of 0.01. The retrieved homologues were then de-replicated with CD-HIT at 70% identity, aligned with Muscle, trimmed with trimAl (--gt 0.1). The resulting sequence alignment was used to reconstruct the maximum-likelihood tree using RAxML v8.2.12. The clustering pattern and decoration of the tree were performed using iTOL v6 (ref. 116).
Environmental distribution of Desulfurivibrionaceae with genomic potential of MISO
The metabolic potential of members belonging to the Desulfurivibrionaceae family was analysed using publicly available genomes recovered from different environments. Metagenome-assembled genomes (MAGs) classified as Desulfurivibrionaceae were obtained from GTDB r214 (n = 121), NCBI (n = 9), JGI IMG (n = 68) and GMGC (n = 7). The environmental origins of these genomes were retrieved from the metadata in the respective databases (Supplementary Table 8). The taxonomy of collected genomes was assigned using GTDB-tk version 2.3.2 with database release 214 (ref. 117). The phylogenomic tree of Desulfurivibrionaceae was reconstructed from a concatenated alignment of 120 single-copy genes with FastTree v2.1.10 (ref. 66). The protein-coding genes in the genomes were called using Prodigal v2.6.3 (ref. 103), and the resulting proteomes were screened for proteins involved in dissimilatory sulfide oxidation (DsrAB) and iron oxides reduction (that is, OmcS, OmcZ, porin–cytochrome complex and OmcE). DsrAB was detected using HMMs and the phylogenetic framework established in this study, while proteins involved in dissimilatory iron reduction were identified with HMMs from FeGenie86. Additional proteins likely involved in dissimilatory reduction of iron oxides—that is, extracellular MHC DA_402 and PilA—were retrieved from Desulfurivibrionaceae genomes by hmmsearch or BLASTP. Homologues of PilA were extracted by searching TIGR02532 HMM model against the Desulfurivibrionaceae genomes using hmmsearch with --cut_ga option. For DA_402, homologues were collected from the Desulfurivibrionaceae genomes using BLASTP with an E value of 1e-10, followed by prediction of the subcellular localization and counting of haem-binding sites. The extracellular homologues containing multi-haem-binding sites (n > 3) were then placed into a reference tree created through phylogenetic analyses of DA_402 (see above) with the RAxML evolutionary placement algorithm (EPA). The alignment for EPA was generated using MAFFT v7.407 with --add option. The homologues that were placed with accumulated probability over 0.95 to the OmcS-like clade were considered as functional orthologs of DA_402. For visualization purposes, Desulfurivibrionaceae genomes (n = 119) encoding both dissimilatory iron and sulfur metabolism were de-replicated on the basis of relative evolutionary divergence (RED). RED was calculated for each internal node of the Desulfurivibrionaceae phylogenomic tree following the procedure described previously118. The tree was then collapsed at a RED value of 0.9 and one representative was chosen randomly from the collapsed clades, yielding 53 representative members that were visualized in the tree.
Statistics and reproducibility
The physiological experiments showing the ability of ferrihydrite-incubated D. alkaliphilus to oxidize formate (Fig. 3a), FeS (Fig. 3b), 1 mM sulfide (Fig. 3c,d,g) or ~50 µM sulfide (Fig. 3h) were repeated independently at least three times, all yielding consistent results. The sulfide removal kinetic experiments at low sulfide concentration were replicated independently for two times, and all results are present in the Extended Data Fig. 1. The experiment showing the transformation of sulfide and ferrihydrite with periodic supply of ~50 µM sulfide was performed independently twice, and both yielded similar results. The growth experiments of ferrihydrite-incubated cells with periodic addition of 1 mM sulfide (Fig. 4a), FeS (Fig. 4b) or 50 µM sulfide (Fig. 4c), and the experiment with nitrate and sulfide (Fig. 4d) were conducted once with three biological replicates per treatment and control. The 13C-bicarbonate labelling experiment and bulk 13C analysis of cells incubated with ferrihydrite and either dissolved sulfide (Fig. 4e) or FeS (Fig. 4f) were independently repeated for two times, both yielding comparable outcomes. Samples for NanoSISM analysis were chosen randomly among biological replicates collected at day 0 and 5, and representative field views are present in Fig. 4g–i. Transcriptomic analysis of cells growing under five different conditions was conducted once, with four biological replicates per condition (Fig. 5 and Extended Data Fig. 4).
Reporting summary
Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.
Data availability
The GTDB r95 genomes were retrieved from the GTDB repository (https://data.ace.uq.edu.au/public/gtdb/data/releases/release95/95.0/genomic_files_reps/). The protein structure of OmcS from G. sulfurreducens was retrieved from the PDB accession 6EF8. Reference sequences and gene-specific HMMs mentioned in the Methods were acquired from SwissProt (https://www.uniprot.org/), KEGG (https://www.kegg.jp/), PFAM (http://pfam.xfam.org/), EggNOG v5.0 (http://eggnog5.embl.de/#/app/home), Sandpiper (https://sandpiper.qut.edu.au/), dbCAN (https://bcb.unl.edu/dbCAN2/download/), metabolicHMM (https://github.com/elizabethmcd/metabolisHMM), CANT-HYD (https://github.com/dgittins/CANT-HYD-HydrocarbonBiodegradation), MicRhoDE (http://application.sb-roscoff.fr/micrhode/), FeGeneie (https://github.com/Arkadiy-Garber/FeGenie), NCBI (https://www.ncbi.nlm.nih.gov/) and JGI IMG (https://img.jgi.doe.gov/). The transcriptomic data of D. alkaliphilus cultured under five conditions (each with four replicates) have been deposited at the NCBI under BioProject PRJNA1165744 (NCBI Sequence Read Archive (SRA) accession: SRX26208148–SRX26208167). HMM models for the 116 sulfur-cycling genes have been deposited at Github (https://github.com/SongCanChen11/HMMs). Source data are provided with this paper.
Code availability
A custom script to retrieve homologues of sulfur-cycling marker genes from GTDB database has been deposited at Github (https://github.com/SongCanChen11/HMMs/).
References
Canfield, D. E. & Farquhar, J. in Fundamentals of Geobiology 49–64 (John Wiley & Sons, 2012).
Fike, D. A., Bradley, A. S. & Rose, C. V. Rethinking the ancient sulfur cycle. Annu. Rev. Earth Planet. Sci. 43, 593–622 (2015).
Wasmund, K., Mußmann, M. & Loy, A. The life sulfuric: microbial ecology of sulfur cycling in marine sediments. Environ. Microbiol. Rep. 9, 323–344 (2017).
Falkowski, P. G., Fenchel, T. & Delong, E. F. The microbial engines that drive Earth’s biogeochemical cycles. Science 320, 1034–1039 (2008).
Mateos, K. et al. The evolution and spread of sulfur cycling enzymes reflect the redox state of the early Earth. Sci. Adv. 9, eade4847 (2023).
Canfield, D. E. & Raiswell, R. The evolution of the sulfur cycle. Am. J. Sci. 299, 697–723 (1999).
Poulton, S. W. Sulfide oxidation and iron dissolution kinetics during the reaction of dissolved sulfide with ferrihydrite. Chem. Geol. 202, 79–94 (2003).
Canfield, D. E. Reactive iron in marine sediments. Geochim. Cosmochim. Acta 53, 619–632 (1989).
Anantharaman, K. et al. Expanded diversity of microbial groups that shape the dissimilatory sulfur cycle. ISME J. 12, 1715–1728 (2018).
Müller, A. L., Kjeldsen, K. U., Rattei, T., Pester, M. & Loy, A. Phylogenetic and environmental diversity of DsrAB-type dissimilatory (bi)sulfite reductases. ISME J. 9, 1152–1165 (2015).
Diao, M. et al. Global diversity and inferred ecophysiology of microorganisms with the potential for dissimilatory sulfate/sulfite reduction. FEMS Microbiol. Rev. 47, fuad058 (2023).
Jørgensen, B. B., Findlay, A. J. & Pellerin, A. The biogeochemical sulfur cycle of marine sediments. Front. Microbiol. 10, 849 (2019).
Canfield, D. E. et al. A cryptic sulfur cycle in oxygen-minimum-zone waters off the Chilean coast. Science 330, 1375–1378 (2010).
Pester, M., Knorr, K.-H., Friedrich, M. W., Wagner, M. & Loy, A. Sulfate-reducing microorganisms in wetlands - fameless actors in carbon cycling and climate change. Front. Microbiol. 3, 72 (2012).
Henkel, J. V. et al. A bacterial isolate from the Black Sea oxidizes sulfide with manganese(IV) oxide. Proc. Natl Acad. Sci. USA 116, 12153–12155 (2019).
Hayes, J. M. & Waldbauer, J. R. The carbon cycle and associated redox processes through time. Philos. Trans. R. Soc. B 361, 931–950 (2006).
Jørgensen, B. B. & Nelson, D. C. in Sulfur Biogeochemistry—Past and Present 63–81 (Geological Society of America, 2004).
Hansel, C. M. et al. Dominance of sulfur-fueled iron oxide reduction in low-sulfate freshwater sediments. ISME J. 9, 2400–2412 (2015).
Flynn, T. M., O’Loughlin, E. J., Mishra, B., DiChristina, T. J. & Kemner, K. M. Sulfur-mediated electron shuttling during bacterial iron reduction. Science 344, 1039–1042 (2014).
Findlay, A. J., Pellerin, A., Laufer, K. & Jørgensen, B. B. Quantification of sulphide oxidation rates in marine sediment. Geochim. Cosmochim. Acta 280, 441–452 (2020).
Michaud, A. B. et al. Glacial influence on the iron and sulfur cycles in Arctic fjord sediments (Svalbard). Geochim. Cosmochim. Acta 280, 423–440 (2020).
Elsgaard, L. & Jørgensen, B. B. Anoxie transformations of radiolabeled hydrogen sulfide in marine and freshwater sediments. Geochim. Cosmochim. Acta 56, 2425–2435 (1992).
Parks, D. H. et al. GTDB: an ongoing census of bacterial and archaeal diversity through a phylogenetically consistent, rank normalized and complete genome-based taxonomy. Nucleic Acids Res. 50, D785–D794 (2022).
Raiswell, R. & Canfield, D. E. The iron biogeochemical cycle past and present. Geochem. Perspect. 1, 1–220 (2012).
Osorio, H. et al. Anaerobic sulfur metabolism coupled to dissimilatory iron reduction in the extremophile Acidithiobacillus ferrooxidans. Appl. Environ. Microbiol. 79, 2172–2181 (2013).
Malik, L. & Hedrich, S. Ferric iron reduction in extreme acidophiles. Front. Microbiol. 12, 818414 (2021).
Thorup, C., Schramm, A., Findlay, A. J., Finster, K. W. & Schreiber, L. Disguised as a sulfate reducer: Growth of the Deltaproteobacterium Desulfurivibrio alkaliphilus by sulfide oxidation with nitrate. mBio 8, e00671-17 (2017).
Wang, F. et al. Structure of microbial nanowires reveals stacked hemes that transport electrons over micrometers. Cell 177, 361–369.e10 (2019).
Shi, L., Fredrickson, J. K. & Zachara, J. M. Genomic analyses of bacterial porin-cytochrome gene clusters. Front. Microbiol. 5, 657 (2014).
Portela, P. C. et al. Widespread extracellular electron transfer pathways for charging microbial cytochrome OmcS nanowires via periplasmic cytochromes PpcABCDE. Nat. Commun. 15, 2434 (2024).
Nübel, T., Klughammer, C., Huber, R., Hauska, G. & Schütz, M. Sulfide:quinone oxidoreductase in membranes of the hyperthermophilic bacterium Aquifex aeolicus (VF5). Arch. Microbiol. 173, 233–244 (2000).
Schütz, M., Maldener, I., Griesbeck, C. & Hauska, G. Sulfide-quinone reductase from Rhodobacter capsulatus: requirement for growth, periplasmic localization, and extension of gene sequence analysis. J. Bacteriol. 181, 6516–6523 (1999).
Finneran, K. T., Johnsen, C. V. & Lovley, D. R. Rhodoferax ferrireducens sp. nov., a psychrotolerant, facultatively anaerobic bacterium that oxidizes acetate with the reduction of Fe(iii). Int. J. Syst. Evol. Microbiol. 53, 669–673 (2003).
Baker, I. R., Conley, B. E., Gralnick, J. A. & Girguis, P. R. Evidence for horizontal and vertical transmission of Mtr-mediated extracellular electron transfer among the bacteria. mBio 13, e02904–e02921 (2021).
Nixon, S. L., Bonsall, E. & Cockell, C. S. Limitations of microbial iron reduction under extreme conditions. FEMS Microbiol. Rev. 46, fuac033 (2022).
Jakobsen, R. & Postma, D. Redox zoning, rates of sulfate reduction and interactions with Fe-reduction and methanogenesis in a shallow sandy aquifer, Rømø, Denmark. Geochim. Cosmochim. Acta 63, 137–151 (1999).
Sorokin, D. Y., Tourova, T. P., Mussmann, M. & Muyzer, G. Dethiobacter alkaliphilus gen. nov. sp. nov., and Desulfurivibrio alkaliphilus gen. nov. sp. nov.: two novel representatives of reductive sulfur cycle from soda lakes. Extremophiles 12, 431–439 (2008).
Schippers, A. & Jørgensen, B. B. Oxidation of pyrite and iron sulfide by manganese dioxide in marine sediments. Geochim. Cosmochim. Acta 65, 915–922 (2001).
Finster, K. Microbiological disproportionation of inorganic sulfur compounds. J. Sulfur Chem. 29, 281–292 (2008).
Sun, J., Wei, L., Yin, R., Jiang, F. & Shang, C. Microbial iron reduction enhances in-situ control of biogenic hydrogen sulfide by FeOOH granules in sediments of polluted urban waters. Water Res. 171, 115453 (2020).
Kjeldsen, K. U. et al. On the evolution and physiology of cable bacteria. Proc. Natl Acad. Sci. USA 116, 19116–19125 (2019).
Santos, T. C., Silva, M. A., Morgado, L., Dantas, J. M. & Salgueiro, C. A. Diving into the redox properties of Geobacter sulfurreducens cytochromes: a model for extracellular electron transfer. Dalton Trans. 44, 9335–9344 (2015).
Kappler, A. et al. An evolving view on biogeochemical cycling of iron. Nat. Rev. Microbiol. 19, 360–374 (2021).
Lovley, D. R. & Holmes, D. E. Electromicrobiology: the ecophysiology of phylogenetically diverse electroactive microorganisms. Nat. Rev. Microbiol. 20, 5–19 (2022).
Filman, D. J. et al. Cryo-EM reveals the structural basis of long-range electron transport in a cytochrome-based bacterial nanowire. Commun. Biol. 2, 219 (2019).
Johs, A., Shi, L., Droubay, T., Ankner, J. F. & Liang, L. Characterization of the decaheme c-type cytochrome OmcA in solution and on hematite surfaces by small angle x-ray scattering and neutron reflectometry. Biophys. J. 98, 3035–3043 (2010).
Antunes, J. M. A., Silva, M. A., Salgueiro, C. A. & Morgado, L. Electron flow from the inner membrane towards the cell exterior in Geobacter sulfurreducens: Biochemical characterization of cytochrome CbcL. Front. Microbiol. 13, 898015 (2022).
Walker, D. J. et al. Electrically conductive pili from pilin genes of phylogenetically diverse microorganisms. ISME J. 12, 48–58 (2018).
Wang, F., Craig, L., Liu, X., Rensing, C. & Egelman, E. H. Microbial nanowires: type IV pili or cytochrome filaments? Trends Microbiol. 31, 384–392 (2023).
Wang, Y. et al. Structural mechanisms of Tad pilus assembly and its interaction with an RNA virus. Sci. Adv. 10, eadl4450 (2024).
Gu, Y. et al. Structure of Geobacter pili reveals secretory rather than nanowire behaviour. Nature 597, 430–434 (2021).
Sonani, R. R. et al. Tad and toxin-coregulated pilus structures reveal unexpected diversity in bacterial type IV pili. Proc. Natl Acad. Sci. USA 120, e2316668120 (2023).
Sun, M. et al. Microbial communities involved in electricity generation from sulfide oxidation in a microbial fuel cell. Biosens. Bioelectron. 26, 470–476 (2010).
de Rink, R. et al. Continuous electron shuttling by sulfide oxidizing bacteria as a novel strategy to produce electric current. J. Hazard. Mater. 424, 127358 (2022).
Ter Heijne, A., de Rink, R., Liu, D., Klok, J. B. M. & Buisman, C. J. N. Bacteria as an electron shuttle for sulfide oxidation. Environ. Sci. Technol. Lett. 5, 495–499 (2018).
Jung, H., Baek, G. & Lee, C. Magnetite-assisted in situ microbial oxidation of H2S to S0 during anaerobic digestion: A new potential for sulfide control. Chem. Eng. J. 397, 124982 (2020).
Jung, H., Yu, H. & Lee, C. Direct interspecies electron transfer enables anaerobic oxidation of sulfide to elemental sulfur coupled with CO2-reducing methanogenesis. iScience 26, 107504 (2023).
Holmkvist, L., Ferdelman, T. G. & Jørgensen, B. B. A cryptic sulfur cycle driven by iron in the methane zone of marine sediment (Aarhus Bay, Denmark). Geochim. Cosmochim. Acta 75, 3581–3599 (2011).
Mikucki, J. A. et al. A contemporary microbially maintained subglacial ferrous ‘ocean’. Science 324, 397–400 (2009).
Och, L. M. et al. New insights into the formation and burial of Fe/Mn accumulations in Lake Baikal sediments. Chem. Geol. 330-331, 244–259 (2012).
Boutet, E., Lieberherr, D., Tognolli, M., Schneider, M. & Bairoch, A. UniProtKB/Swiss-Prot. Methods Mol. Biol. 406, 89–112 (2007).
Fu, L., Niu, B., Zhu, Z., Wu, S. & Li, W. CD-HIT: Accelerated for clustering the next-generation sequencing data. Bioinformatics 28, 3150–3152 (2012).
Camacho, C. et al. BLAST+: architecture and applications. BMC Bioinformatics 10, 421 (2009).
Edgar, R. C. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 32, 1792–1797 (2004).
Capella-Gutiérrez, S., Silla-Martínez, J. M. & Gabaldón, T. trimAl: A tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25, 1972–1973 (2009).
Price, M. N., Dehal, P. S. & Arkin, A. P. FastTree 2—approximately maximum-likelihood trees for large alignments. PLoS ONE 5, e9490 (2010).
Marcia, M., Ermler, U., Peng, G. & Michel, H. A new structure-based classification of sulfide:quinone oxidoreductases. Proteins 78, 1073–1083 (2010).
Orellana, L. H., Rodriguez-R, L. M. & Konstantinidis, K. T. ROCker: Accurate detection and quantification of target genes in short-read metagenomic data sets by modeling sliding-window bitscores. Nucleic Acids Res. 45, e14 (2017).
Eddy, S. R., Wheeler, T. J. & the HMMER development team. HMMER User’s Guide Version 3.1b2, 108 (Howard Hughes Medical Institute, 2015); http://eddylab.org/software/hmmer3/3.1b2/Userguide.pdf.
Aramaki, T. et al. KofamKOALA: KEGG Ortholog assignment based on profile HMM and adaptive score threshold. Bioinformatics 36, 2251–2252 (2020).
Haft, D. H. et al. TIGRFAMs: a protein family resource for the functional identification of proteins. Nucleic Acids Res. 29, 41–43 (2001).
Finn, R. D. et al. Pfam: the protein families database. Nucleic Acids Res. 42, D222–D230 (2014).
McDaniel, E. A., Anantharaman, K. & McMahon, K. D. metabolisHMM: Phylogenomic analysis for exploration of microbial phylogenies and metabolic pathways. Preprint at bioRxiv https://doi.org/10.1101/2019.12.20.884627 (2019).
Neukirchen, S. & Sousa, F. L. DiSCo: a sequence-based type-specific predictor of Dsr-dependent dissimilatory sulphur metabolism in microbial data. Microb. Genomics 7, 000603 (2021).
Teng, Z.-J. et al. Biogeographic traits of dimethyl sulfide and dimethylsulfoniopropionate cycling in polar oceans. Microbiome 9, 207 (2021).
Tanabe, T. S. & Dahl, C. HMS-S-S: A tool for the identification of sulphur metabolism-related genes and analysis of operon structures in genome and metagenome assemblies. Mol. Ecol. Resour. 22, 2758–2774 (2022).
Yu, G., Smith, D. K., Zhu, H., Guan, Y. & Lam, T. T.-Y. ggtree: an R package for visualization and annotation of phylogenetic trees with their covariates and other associated data. Methods Ecol. Evol. 8, 28–36 (2017).
Gupta, D., Chen, K., Elliott, S. J. & Nayak, D. D. MmcA is an electron conduit that facilitates both intracellular and extracellular electron transport in Methanosarcina acetivorans. Nat. Commun. 15, 3300 (2024).
Leu, A. O. et al. Lateral gene transfer drives metabolic flexibility in the anaerobic methane-oxidizing archaeal family Methanoperedenaceae. mBio 11, e01325 (2020).
Garber, A. I., Nealson, K. H. & Merino, N. Large-scale prediction of outer-membrane multiheme cytochromes uncovers hidden diversity of electroactive bacteria and underlying pathways. Front. Microbiol. 15, 1448685 (2024).
Moreno, J., Nielsen, H., Winther, O. & Teufel, F. Predicting the subcellular location of prokaryotic proteins with DeepLocPro. Bioinformatics 40, btae677 (2024).
Huerta-Cepas, J. et al. eggNOG 5.0: A hierarchical, functionally and phylogenetically annotated orthology resource based on 5090 organisms and 2502 viruses. Nucleic Acids Res. 47, D309–D314 (2019).
Yin, Y. et al. dbCAN: A web resource for automated carbohydrate-active enzyme annotation. Nucleic Acids Res. 40, W445–W451 (2012).
Khot, V. et al. CANT-HYD: A curated database of phylogeny-derived Hidden Markov Models for annotation of marker genes involved in hydrocarbon degradation. Front. Microbiol. 12, 764058 (2021).
Boeuf, D., Audic, S., Brillet-Guéguen, L., Caron, C. & Jeanthon, C. MicRhoDE: A curated database for the analysis of microbial rhodopsin diversity and evolution. Database 2015, bav080 (2015).
Garber, A. I. et al. FeGenie: A comprehensive tool for the identification of iron genes and iron gene neighborhoods in genome and metagenome assemblies. Front. Microbiol. 11, 37 (2020).
Teufel, F. et al. SignalP 6.0 predicts all five types of signal peptides using protein language models. Nat. Biotechnol. 40, 1023–1025 (2022).
Woodcroft, B. J. et al. Comprehensive taxonomic identification of microbial species in metagenomic data using SingleM and Sandpiper. Nat. Biotechnol. https://doi.org/10.1038/s41587-025-02738-1 (2025).
Amend, J. P. & LaRowe, D. E. Minireview: Demystifying microbial reaction energetics. Environ. Microbiol. 21, 3539–3547 (2019).
Cornell, R. M. & Schwertmann, U. The Iron Oxides: Structure, Properties, Reactions, Occurrence and Uses (Wiley-VCH, 2003).
Rickard, D. & Morse, J. W. Acid volatile sulfide (AVS). Mar. Chem. 97, 141–197 (2005).
Majzlan, J., Navrotsky, A. & Schwertmann, U. Thermodynamics of iron oxides: part III. Enthalpies of formation and stability of ferrihydrite (∼Fe(OH)3), schwertmannite (∼FeO(OH)3/4(SO4)1/8), and ε-Fe2O3. Geochim. Cosmochim. Acta 68, 1049–1059 (2004).
Ritz, C., Baty, F., Streibig, J. C. & Gerhard, D. Dose–response analysis using R. PLoS ONE 10, e0146021 (2015).
Poulton, S. W., Krom, M. D. & Raiswell, R. A revised scheme for the reactivity of iron (oxyhydr)oxide minerals towards dissolved sulfide. Geochim. Cosmochim. Acta 68, 3703–3715 (2004).
Cline, J. D. Spectrophotometric determination of hydrogen sulfide in natural waters. Limnol. Oceanogr. 14, 454–458 (1969).
Wan, M., Shchukarev, A., Lohmayer, R., Planer-Friedrich, B. & Peiffer, S. Occurrence of surface polysulfides during the interaction between ferric (hydr)oxides and aqueous sulfide. Environ. Sci. Technol. 48, 5076–5084 (2014).
Ehmann, T. et al. Modification and validation of the pyromellitic acid electrolyte for the capillary electrophoretic determination of anions. J. Chromatogr. A 995, 217–226 (2003).
Thamdrup, B., Finster, K., Hansen, J. W. & Bak, F. Bacterial disproportionation of elemental sulfur coupled to chemical reduction of iron or manganese. Appl. Environ. Microbiol. 59, 101–108 (1993).
Hausmann, B. et al. Consortia of low-abundance bacteria drive sulfate reduction-dependent degradation of fermentation products in peat soil microcosms. ISME J. 10, 2365–2375 (2016).
Griffiths, R. I., Whiteley, A. S., O’Donnell, A. G. & Bailey, M. J. Rapid method for coextraction of DNA and RNA from natural environments for analysis of ribosomal DNA- and rRNA-based microbial community composition. Appl. Environ. Microbiol. 66, 5488–5491 (2000).
Ewels, P., Magnusson, M., Lundin, S. & Käller, M. MultiQC: summarize analysis results for multiple tools and samples in a single report. Bioinformatics 32, 3047–3048 (2016).
Liao, Y., Smyth, G. K. & Shi, W. The Subread aligner: fast, accurate and scalable read mapping by seed-and-vote. Nucleic Acids Res. 41, e108 (2013).
Hyatt, D. et al. Prodigal: prokaryotic gene recognition and translation initiation site identification. BMC Bioinformatics 11, 119 (2010).
Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).
Haynes, W. in Encyclopedia of Systems Biology (eds Dubitzky, W. et al.) 78–78 (Springer, 2013).
Ye, J. et al. Primer-BLAST: a tool to design target-specific primers for polymerase chain reaction. BMC Bioinformatics 13, 134 (2012).
Deng, X., Dohmae, N., Nealson, K. H., Hashimoto, K. & Okamoto, A. Multi-heme cytochromes provide a pathway for survival in energy-limited environments. Sci. Adv. 4, eaao5682 (2018).
Yu, N. Y. et al. PSORTb 3.0: Improved protein subcellular localization prediction with refined localization subcategories and predictive capabilities for all prokaryotes. Bioinformatics 26, 1608–1615 (2010).
Almagro Armenteros, J. J. et al. SignalP 5.0 improves signal peptide predictions using deep neural networks. Nat. Biotechnol. 37, 420–423 (2019).
Notredame, C., Higgins, D. G. & Heringa, J. T-Coffee: a novel method for fast and accurate multiple sequence alignment. J. Mol. Biol. 302, 205–217 (2000).
Holm, L. & Park, J. DaliLite workbench for protein structure comparison. Bioinformatics 16, 566–567 (2000).
Holm, L. Benchmarking fold detection by DaliLite v.5. Bioinformatics 35, 5326–5327 (2019).
Zhang, Y. & Skolnick, J. Scoring function for automated assessment of protein structure template quality. Proteins 57, 702–710 (2004).
Morris, G. M. et al. AutoDock4 and AutoDockTools4: automated docking with selective receptor flexibility. J. Comput. Chem. 30, 2785–2791 (2009).
Eberhardt, J., Santos-Martins, D., Tillack, A. F. & Forli, S. AutoDock Vina 1.2.0: new docking methods, expanded force field, and Python bindings. J. Chem. Inf. Model. 61, 3891–3898 (2021).
Letunic, I. & Bork, P. Interactive Tree of Life (iTOL) v6: Recent updates to the phylogenetic tree display and annotation tool. Nucleic Acids Res. 52, W78–W82 (2024).
Chaumeil, P.-A., Mussig, A. J., Hugenholtz, P. & Parks, D. H. GTDB-Tk: A toolkit to classify genomes with the Genome Taxonomy Database. Bioinformatics 36, 1925–1927 (2019).
Parks, D. H. et al. A standardized bacterial taxonomy based on genome phylogeny substantially revises the tree of life. Nat. Biotechnol. 36, 996–1004 (2018).
Cai, L. et al. Tad pilus-mediated twitching motility is essential for DNA uptake and survival of Liberibacters. PLoS ONE 16, e0258583 (2021).
Melville, S., & Craig, L. Type IV pili in Gram-positive bacteria. Microbiol. Mol. Biol. Rev. 77, 323–341 (2013).
Acknowledgements
The authors thank B. Hausmann, K. Wasmund and T. S. Tanabe for help with selection of sulfur-cycling target genes; M. Schmid for Raman analyses of ferrihydrite; J. Ramesmayer for preparation of cDNA libraries; L. Seidl for technical assistance with elemental sulfur analysis; M. Watzka for isotope ratio mass spectrometry analysis; N. Kumar for support in ferrihydrite synthesis; O. Coruh for help with protein structure modelling; K. Schmidt, S. Dürr and D. Gruber for SEM and TEM imaging; T. Rattei and his team for maintaining and providing access to the Life Science Compute Cluster (https://lisc.univie.ac.at/); the Bundesministerium für Bildung und Forschung (BMBF)-funded deNBI cloud within German Network for Bioinformatics Infrastructure (de.NBI) (no. 031A532B, 031A533A, 031A533B, 031A534A, 031A535A, 031A537A, 031A537B, 031A537C, 031A537D and 031A538A) for providing computational resources; and H. Daims for valuable discussion. The drawing of pili was inspired by figures in Cai et al. (2021)119 and Melville & Craig (2013)120. This research was funded by the Austrian Science Fund (FWF) (grants https://doi.org/10.55776/P31996 and https://doi.org/10.55776/COE7), the EU MSCA postdoctoral fellowship (action number 101059607, DatingSuCy) to S.-C.C., the National Natural Science Foundation of China (grants 42307167 and 42021005), and a Chinese Scholarship Council fellowship to X.-M.L. For open access purposes, the author has applied a CC BY public copyright license to any author accepted manuscript version arising from this submission.
Funding
Open access funding provided by University of Vienna.
Author information
Authors and Affiliations
Contributions
S.-C.C., M.M. and A.L. conceived the study and planned experiments, with help from P.P. and S.P. S.-C.C. performed bioinformatic analyses. S.-C.C., X.-M.L., N.B., G.G. and M.A.M. designed and performed physiological experiments. S.-C.C., N.B., M.M. and M.A.M. quantified cell growth and metabolite concentrations. W.W. contributed to elemental sulfur and isotope ratio mass spectrometry analysis. A.S. performed NanoSIMS analysis. S.-C.C. and M.A.M. extracted RNA from cultures incubated under different conditions. S.-C.C., J.O. and P.P. performed comparative transcriptome analyses. W.W. and A.R. provided essential infrastructure for metabolite and isotope analyses. S.-C.C. analysed the data. S.-C.C., M.M. and A.L. interpreted the data. S.-C.C. and A.L. wrote the paper, with comments from all other authors.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature thanks Andreas Schramm and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer review reports are available.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Extended data figures and tables
Extended Data Fig. 1 Sulfide removal rates at low concentration of sulfide (ca. 50 µM) are higher in ferrihydrite-amended D. alkaliphilus cultures than in abiotic controls.
a, Kinetics of dissolved sulfide during 25-min incubation with ferrihydrite (ca. 62 mM) in the presence or absence of D. alkaliphilus. The incubation experiment was conducted independently with three different cell densities, indicated by optical density at 600 nm (OD600) of the inoculum. Each incubation received three spikes of sulfide at 1.5-h intervals. Two replicates were performed for each condition. OD600 of 0.042 corresponds to an average cell density of 1.38 × 108 cells ml−1. b, Sulfide concentrations in the sulfide-only control of the incubation experiments remained constant. The residual sulfide carried over from the inoculum was subtracted from the measured sulfide concentration. c, The rate constant of sulfide consumption was significantly higher (P < 0.01; two-sided Student’s t-test) in D. alkaliphilus cultures than in abiotic controls. The rate was determined by fitting the first-order rate equation to the measured sulfide concentration changes (see panel a). The center lines and box limits of the boxplot denote the median, and the 25% and 75% percentile of the estimated rate constant, respectively. The whiskers extend 1.5 times the interquartile range from the 25th and 75th percentiles. The two-sided Student’s t-test between biotic (n = 6) and abiotic (n = 6) treatment was conducted for each cell density. P values were adjusted for multiple comparisons (q-values) using the Benjamini and Hochberg method. d, Concentration of HCl-extractable Fe(II) before three consecutive spikes of sulfide to ferrihydrite with or without D. alkaliphilus. The values represent the mean of duplicate culture with OD600 of 0.086. Fe(II) concentration in both treatments increased with repeated spikes of low conditions of sulfide. The treatment with cells produced an excessive amount of Fe(II) compared to the abiotic control. Fe(II) before the 1st sulfide spike results from biological or chemical oxidation of sulfide (ca. 50 µM) carried over from the inoculum.
Extended Data Fig. 2 Transformation of sulfide during the incubation of D. alkaliphilus with ferrihydrite (ca. 62 mM) and periodic supply (n = 8) of a low concentration of sulfide (ca. 50 µM) at 1.5-h intervals.
a, Development of sulfate, HCl-extractable Fe(II), elemental sulfur (S(0)), and Cline-extractable sulfide during the incubation experiment (n = 3 replicate cultures). Parallel incubations with living cells without Fh or sulfide, and with killed cells served as controls. The arrows in each panel indicate the addition of sulfide. The red stars denote sampling from the culture for chemical analysis. The inoculum used for the incubation experiment had an OD600 of 0.072. b, Kinetics of dissolved sulfide following the first three spikes to ferrihydrite-amended culture in a parallel incubation experiment with the same inoculum as in panel a. Sulfide measurements from two biological replicates are displayed. The data confirmed that the culture consumed sulfide more rapidly than the abiotic control. c, Schematic graph showing the flow of spiked sulfide during the 12-h incubation. Sulfur metabolite concentrations derive from data in panel a. All values are the means of triplicate cultures with standard deviation.
Extended Data Fig. 3 Genomic arrangement of genes putatively involved in MISO.
a, Genes for reversal of the canonical sulfate reduction pathway. b, Genes for the up-regulated multi-heme c-type cytochromes (MHCs) during MISO. c, Genes for the PilA-containing type IV pilus. d, Genes for the tight adhesion (Tad) type IV pilus. The gene name and locus tag are shown above and below the gene arrows, respectively. Count of heme-binding motifs (CX2-5CH) is shown for each MHC in b. Co-localized genes not shown in Fig. 5 are colored in gray.
Extended Data Fig. 4 Up-regulated transcription of 13 multi-heme c-type cytochromes (MHCs) in D. alkaliphilus under ferrihydrite-amended conditions.
Comparative transcriptomics of three ferrihydrite-amended and two ferrihydrite-free growth conditions (n = 4 replicate cultures each). The heatmap is based on the z-score of reads per kilobase of transcript per million mapped reads (RPKM) after logarithm transformation. The statistical significance (P-value) of up-regulation was calculated using the DESeq2 R package, followed by adjustment of P for multiple comparison using the Benjamini and Hochberg method. The 13 MHCs showed significantly elevated relative transcription levels (adjusted P < 0.01) in all pairwise comparisons (n = 6) between ferrihydrite-amended incubations (n = 3) vs. incubations (n = 2) without ferrihydrite. Electron donors and acceptors for each of the five incubations are indicated above the heatmap. Four biological replicates of each incubation are displayed side-by-side. The subcellular localization of MHCs was predicted by Psortb v3.0 and DeepLocPro v1.0. PCC MHC, MHC from a gene cluster encoding a putative porin-cytochrome complex (PCC).
Extended Data Fig. 5 Sequence alignment and predicted structure of the most expressed multi-heme c-type cytochrome (DA_402) of Desulfurivibrio alkaliphilus during MISO.
a, Pairwise alignment between DA_402 and the OmcS from G. sulfurreducens. Six out of seven predicted heme-binding motifs (CX2-5CH) are aligned with those in OmcS (labeled with numbers in circles). The seventh heme-binding motif lacks the paired histidine (blue color) for coordination of heme in the predicted structure of DA_402 and is therefore not considered in the following analysis. The leading signal peptide in DA_402 and OmcS is not shown. b, Predicted structure of DA_402 oligomer. DA_402 monomers are predicted to polymerize into a filament structure, similar to the cryo-EM structure of OmcS28. c, Structural comparison between DA_402 and OmcS. DA_402 shares significant topological similarity with OmcS, with a TM-score of 0.56 and DALI-score of 14.3. d, Spatial arrangement of six heme-binding sites in DA_402 (cyan). All heme-binding sites predicted in DA_402 align closely with those in OmcS (yellow), except the site 1’ showing shifted position. e, Predicted heme position of the heme-binding site 1’. Molecular docking analysis placed the heme (1’) in close proximity to the adjacent heme ②, with an iron-to-iron distance of 9.1 Å, comparable with the distance (12.5 Å) observed in OmcS.
Extended Data Fig. 6 Desulfurivibrionaceae members with genomic potential for MISO are widespread across various ecosystems.
a, Co-occurrence of marker proteins for dissimilatory sulfide oxidation and iron reduction in diverse Desulfurivibrionaceae members recovered from different ecosystems. Phylogenomic tree of Desulfurivibrionaceae is based on maximum-likelihood analysis of a concatenated alignment of 120 bacterial single-copy proteins. 119 of 205 Desulfurivibrionaceae genomes encode both dissimilatory sulfur and iron metabolisms. For visualization purposes, the phylogeny includes only 53 de-replicated genomes. D. alkaliphilus is shown in bold and red type. The source ecosystem of each genome is denoted by filled circles at the tips of the tree. Environmental origin, NCBI accession, and genus classification is shown at each branch tip. Desulfobulbaceae members were used as an outgroup. Presence and absence of marker proteins involved in sulfide oxidation (DsrAB, red squares) and extracellular respiration of iron oxides (green squares) is shown for each genome. Abbreviation: OmcS and OmcZ, two evolutionarily unrelated outer membrane cytochromes; PilA, major component (pilin) of the type IV pilus; PCC, porin-cytochrome complex. b, Phylogeny of DsrB from Desulfurivibrionaceae shown in a, which are classified as the reductive-type DsrB putatively involved in sulfide oxidation, as proposed in D. alkaliphilus and in cable bacteria. Known sulfide oxidizers (e.g., cable bacteria genus Electrothrix) encoding reductive-type DsrBs are indicated by red squares. c, Phylogeny of OmcS-like proteins from Desulfurivibrionaceae shown in a, which form a sister clade of OmcS. The sister clade of OmcS comprise extracellular MHCs from known iron reducers (dark green squares) and three MHCs (DA_402, DA_756, and DA_765) from D. alkaliphilus that are up-regulated under MISO conditions. All OmcS-like proteins shown in a are predicted to be extracellular multiheme c-type cytochromes (MHCs). Blue circles indicate the numbers of Desulfurivibrionaceae MHC genes that are placed to the tree branches using RAxML EPA algorithm. The scale bar in a, b, and c indicates the substitutions of amino acids per site. The branches with bootstrap values > 50 are shown with black dots in a and b.
Supplementary information
Supplementary Information
Supplementary text including Supplementary Figs. 1–15 and references.
Supplementary Tables
Supplementary Tables 1–8.
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
About this article
Cite this article
Chen, SC., Li, XM., Battisti, N. et al. Microbial iron oxide respiration coupled to sulfide oxidation. Nature (2025). https://doi.org/10.1038/s41586-025-09467-0
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41586-025-09467-0