Abstract Anoxygenic phototrophic sulfur bacteria flourish in contemporary and ancient euxinic environments, driving the biogeochemical cycles of carbon and sulfur. However, it is unclear how these strict anaerobes meet their high demand for iron in iron-depleted environments. Here, we report that pyrite, a widespread and highly stable iron sulfide mineral in anoxic, low-temperature environments, can support the growth and metabolic activity of anoxygenic phototrophic sulfur bacteria by serving as the sole iron source under iron-depleted conditions. Transcriptomic and proteomic analyses revealed that pyrite addition substantially up-regulated genes and protein expression involved in photosynthesis, sulfur metabolism, and biosynthesis of organics. Anoxic microbial oxidation of pyritic sulfur and consequent destabilization of the pyrite structure were postulated to facilitate microbial iron acquisition. These findings advance our understanding of the survival strategies of anaerobes in iron-depleted environments and are important for revealing the previously underappreciated bioavailability of pyritic iron in anoxic environments and anoxic weathering of pyrite. __________________________________________________________________ Anoxygenic phototrophic sulfur bacteria acquire iron from pyrite under iron-depleted conditions via oxidizing pyritic sulfur. INTRODUCTION Oceanic euxinia (anoxic and sulfidic conditions) ubiquitously occurred in paleo-ocean, particularly during the Proterozoic ([36]1–[37]3). Because of the high toxicity of hydrogen sulfide (H[2]S), euxinia are believed to be an important reason for biotic crises such as the largest known Phanerozoic extinction—the end-Permian mass extinction (~252 million years ago) ([38]4–[39]6). Seasonal euxinia also frequently develop in modern environments such as the Black Sea, fjords, meromictic lakes, and the Benguela Upwelling System ([40]5, [41]7–[42]9). Although euxinia are well known for containing highly toxic H[2]S, with a concentration varying from several hundreds to thousands of micromolar, such hostile environments generally accommodate unique microbial communities such as anoxygenic phototrophic purple or green sulfur bacteria (PSB or GSB) ([43]5, [44]6). These bacteria can produce long-lasting organic biomarkers, such as carotenoid isorenieratene, which can be used for robust identification of photic zone euxinia in ancient environments ([45]4, [46]10, [47]11). Prosperous growth of anoxygenic phototrophic sulfur bacteria requires illumination, carbon source, and reduced sulfur compounds ([48]12). As research progresses regarding the reaction center of anoxygenic photosynthesis, the importance of trace metals, particularly iron, in the growth of anoxygenic phototrophic sulfur bacteria has been revealed recently ([49]12, [50]13). For instance, three key Fe[4]S[4] clusters were identified within the electron transfer chain of the photosynthetic reaction center in the representative strain of GSB Chlorobaculum tepidum ([51]12). During anoxygenic photosynthesis, light energy captured by chlorosome (i.e., the light-harvesting structure of C. tepidum) is transferred by Fenna-Matthews-Olson proteins to its membrane-embedded photosynthetic reaction center to initiate charge separation and electron transfer reactions ([52]12, [53]13). A similarly pivotal role of iron in the growth and carbon fixation of phytoplankton has also been reported ([54]14, [55]15). In addition, iron acts as a crucial cofactor for key enzymes involved in the sulfur oxidation processes of C. tepidum ([56]16). Therefore, fluctuations in iron concentrations over geological timescale are often closely linked to changes in ocean primary productivity and biogeochemical cycles of many elements such as carbon and sulfur ([57]15, [58]17). However, euxinic environments are iron depleted and usually characterized by a high degree of pyritization [e.g., reactive aqueous iron or iron (hydro)oxides react with free H[2]S to form fine-grained pyrite in the water column] ([59]18–[60]20). This creates an apparent paradox between iron scarcity and widespread distribution of GSB-derived biomarkers (i.e., requiring iron for GSB growth) in both the geological past and modern euxinic environments ([61]8, [62]10), prompting inquiries about the iron sources and the survival strategies of GSB in sulfidic settings. Given that pyrite is considered as the most abundant iron form in euxinic environments ([63]2, [64]20), it may be possible that pyrite-associated iron supports the growth of anoxygenic phototrophic sulfur bacteria, despite the usual belief that pyrite is stable under anoxic, low-temperature conditions ([65]21). It is long believed that pyrite is not bioavailable in the absence of oxygen or alternative electron acceptors such as nitrate and ferric iron ([66]22–[67]24). Nevertheless, such belief is challenged by recent lines of evidence of the utilization of sulfur within pyrite by anaerobic PSB and the utilization of both iron and sulfur within pyrite by anaerobic methanogenic archaea ([68]21, [69]25, [70]26). Although the underlying mechanism remains unclear, these studies suggest that anaerobic weathering of pyrite might be possible in the environment, especially under iron- or sulfur-limited conditions. Thus, we speculate that pyrite may serve as an efficient iron source to promote the growth of anoxygenic phototrophic sulfur bacteria. The primary objectives of this study, therefore, were to investigate whether anoxygenic phototrophic sulfur bacteria, using GSB C. tepidum as a representative bacterium, can use pyrite as the sole iron source under iron-depleted conditions, and if so, what the potential underlying mechanisms are. Understanding the metabolic flexibility and adaptation strategies of anoxygenic phototrophic sulfur bacteria in response to iron deficiency will shed light on microbial weathering of pyrite under anoxic condition, which is of great significance to the biogeochemical cycles of sulfur, carbon, and iron in euxinic environments ([71]27). RESULTS Inhibited cell growth and sulfur oxidation under iron-depleted conditions Cultivation experiments showed that C. tepidum grew steadily in the canonical medium regardless of cell transfer cycles, reaching an optical density at 750 nm (OD[750]) value of around 1.7 at the plateau stage (fig. S1). In contrast, the growth of C. tepidum in the iron-depleted medium was increasingly inhibited after multiple transfers, and the OD[750] decreased from 0.9 after the first transfer to 0.5 after the third transfer (fig. S1). Although the growth of C. tepidum was not completely suppressed under iron-depleted conditions, probably because of a trace amount of iron derived from chemical impurities, these results suggested that iron deficiency limited the growth of C. tepidum. When cells collected from the third transfer in the iron-depleted medium were inoculated into a modified medium amended with aqueous Fe^2+ solution (concentration: 0.2, 2, or 20 μM), the cell growth of C. tepidum was restored ([72]Fig. 1A). For instance, the addition of 2 and 20 μM aqueous Fe^2+ notably enhanced the growth of C. tepidum, as indicated by the greater values of OD[750] ([73]Fig. 1A) and higher concentrations of total protein (fig. S2A) and bacteriochlorophyll c (BChl c) (fig. S2B), compared with that of the iron-depleted treatment, with a higher promoting effect with addition of 20 μM Fe^2+ than with 2 μM Fe^2+ ([74]Fig. 1A and fig. S2). However, there was little effect on cell growth when 0.2 μM Fe^2+ was amended ([75]Fig. 1A and fig. S2), suggesting that an iron concentration higher than 0.2 μM was required to sustain the active metabolism of C. tepidum. Fig. 1. Growth curves and time-course changes of sulfide concentration. [76]Fig. 1. [77]Open in a new tab (A) Growth curves of C. tepidum incubated under Fe-depleted conditions or with the addition of various concentrations of aqueous Fe^2+ (0.2, 2, and 20 μM) and solid-phase iron (i.e., pyrite). (B) The corresponding time-course decreases of sulfide concentration. Accompanied by the inhibited cell growth of C. tepidum, microbial sulfur oxidation rate decreased obviously under iron-depleted conditions ([78]Fig. 1B). In general, C. tepidum oxidizes sulfide to extracellular elemental sulfur globules, and subsequent oxidation of elemental sulfur and thiosulfate only occurs when sulfide is depleted ([79]28–[80]30). Such sequential sulfur oxidation was observed in this study ([81]Fig. 1B and fig. S3). For instance, sulfide concentration decreased slowly in the treatments with no or 0.2 μM Fe^2+ ([82]Fig. 1B), concomitant with a gradual accumulation of elemental sulfur (fig. S3A). Neither consumption of thiosulfate nor production of sulfate was observed in these two treatments (fig. S3, B and C) because of the remaining sulfide in the medium ([83]Fig. 1B). In comparison, active sulfur oxidation proceeded in the presence of 2 or 20 μM Fe^2+ as indicated by complete consumption of sulfide after 30 and 36 hours of inoculation, respectively ([84]Fig. 1B). Once sulfide was depleted, elemental sulfur (produced from sulfide oxidation) and initially added thiosulfate began to decrease (fig. S3, A and B), accompanied by sulfate production (fig. S3C). The calculated electron transfer rate within the logarithmic growth phase (i.e., 12 to 36 hours) increased from about 0.04 mM/hour under iron-depleted conditions to 0.45 mM/hour in the presence of 2 μM Fe^2+ and further to 0.77 mM/hour with 20 μM Fe^2+ addition (fig. S4). Enhanced cell growth and sulfur oxidation with pyrite as the sole iron source Results of both OD[750] and total protein concentration showed that the addition of pyrite (1 g/liter) stimulated the growth of C. tepidum irrespective of particle size of pyrite (fig. S5A), with an enhancement effect even slightly better than that with 20 μM aqueous Fe^2+ addition ([85]Fig. 1A and fig. S2). This slightly higher promotion of pyrite suggests that in addition to serving as an iron source, pyrite might stimulate the growth of C. tepidum by providing photoelectrons due to its semiconductive property. Such extracellular photoelectrons have been shown to be used by anoxygenic phototrophs (e.g., PSB Rhodopseudomonas palustris) in CO[2] fixation to biomass ([86]31). However, the promoting effect did not occur when pyrite was replaced with similarly sized quartz (fig. S5A), thus excluding the possibility that the provision of physical attachment sites by solid mineral was responsible for the enhanced cell growth. A more likely reason was the lack of bioessential elements such as iron in quartz ([87]32). Similar to the growth pattern, pyrite accelerated sulfur oxidation, while quartz did not ([88]Fig. 1B and fig. S5B). In the pyrite treatment, sulfide was completely consumed in approximately 30 hours ([89]Fig. 1B), in conjunction with a rapid accumulation of elemental sulfur, which, along with thiosulfate, was subsequently oxidized to sulfate (fig. S3). In contrast, about 25% of the initially added sulfide remained by the end of the reaction (72 hours) in the treatment with quartz (fig. S5B), and no sulfate production was detected. This pattern was similar to the iron-depleted treatment (fig. S5). Moreover, the electron transfer rates during the logarithmic growth period increased from 0.04 mM/hour under iron-depleted conditions to 0.81 mM/hour in the presence of pyrite (fig. S4), indicating that pyrite served as the sole iron source to sustain active microbial sulfur oxidation. Cultivation of C. tepidum in pyrite leachate (i.e., soaking pyrite in the medium for 7 days before inoculation) showed similar growth and sulfide-oxidation kinetics to those in the iron-depleted medium (fig. S6). This result was expected because pyrite is stable under euxinic conditions ([90]20) and led to a negligible amount of iron release into the leachate (i.e., ~0.10 μM), which was not sufficient to meet the cellular iron demand of C. tepidum ([91]Fig. 1). Thus, abiotic dissolution of pyrite could not account for the growth promotion of C. tepidum. In addition, when pyrite was sequestered in a dialysis bag with a molecular weight cutoff of 500 Da, the accelerating effects on both growth and sulfur-oxidation process of C. tepidum were still observed but were much weaker than those without a dialysis bag (fig. S7). These results indicated that either direct contact with pyrite or high–molecular weight (>500 Da) extracellular proteins and metabolites were essential for the microbial acquisition of iron from pyrite. Cellular iron uptake during cell growth In the iron-depleted and 0.2 μM Fe^2+ treatments, both aqueous iron ([92]Fig. 2A) and cell-associated iron ([93]Fig. 2B) were hardly detectable over the course of the experiments. In contrast, aqueous iron concentration in the 20 μM Fe^2+ treatment declined gradually ([94]Fig. 2A), while the cell-associated iron concentration increased over time and reached a plateau after 36 hours ([95]Fig. 2B), in accordance with the growth profile of C. tepidum ([96]Fig. 1A), suggesting substantial cellular uptake of iron during the growth of C. tepidum. In the 2 μM Fe^2+ treatment, aqueous iron in the medium declined near the beginning ([97]Fig. 2A), and a lower accumulation of cell-associated iron was observed ([98]Fig. 2B), in alignment with the weaker growth enhancement of C. tepidum in this treatment ([99]Fig. 1A). Under the same Fe concentration, the biomass-normalized cell-associated iron concentration did not change obviously over time for all treatments ([100]Fig. 2C), which might reflect the rapid metabolic response of C. tepidum to iron availability and a dependence on the bacterial physiological needs ([101]33). The lower concentration of cell-associated iron per unit OD in the 2 μM Fe^2+ treatment than that in the 20 μM Fe^2+ treatment ([102]Fig. 2C) suggested that the optimal iron concentration required by C. tepidum was greater than 2 μM. Fig. 2. Fe concentrations. [103]Fig. 2. [104]Open in a new tab Time-course concentration changes of iron (A) in culture supernatant, (B) in association with cells, and (C) in cell-associated iron per unit OD under Fe-depleted conditions or with the addition of various concentrations of aqueous Fe^2+ (0.2, 2, and 20 μM) and solid-phase iron (i.e., pyrite). In the pyrite treatment, aqueous iron in the medium was hardly detectable over the whole reaction period ([105]Fig. 2A), due to the high stability of pyrite in anoxic environments ([106]34). However, the cell-associated iron concentration increased substantially, even slightly higher than that in the 20 μM Fe^2+ treatment by 72 hours ([107]Fig. 2B). Similar to the aqueous Fe^2+ treatments, the biomass-normalized cell-associated Fe concentration did not change over time ([108]Fig. 2C). Dialysis experiments showed that both absolute and biomass-normalized concentrations of cell-associated iron were substantially diminished when pyrite particles and cells were physically separated (fig. S8). Overall, these data demonstrated that cellular iron uptake by C. tepidum did occur when pyrite was used as the sole, solid-phase iron source, which was responsible for the enhanced cell growth and sulfur oxidation ([109]Fig. 1). Pyrite surface modification by C. tepidum Scanning electron microscopy (SEM) observation displayed a close physical association between C. tepidum cells and pyrite particles ([110]Fig. 3), verifying the importance of direct contact for microbial iron acquisition (figs. S7 and S8). Although x-ray diffraction (XRD) analysis showed no bulk mineralogical change of pyrite or the formation of secondary minerals (fig. S9), the variations in the relative peak intensities between the untreated and treated pyrite suggested that C. tepidum may have modified the surface structure of pyrite to facilitate iron acquisition. Meanwhile, time-of-flight secondary ion mass spectrometry (TOF-SIMS) analysis revealed chemical modification (i.e., elemental composition changes) of pyrite surface after its interaction with C. tepidum under iron-depleted conditions ([111]Fig. 4). Specifically, principal components analysis (PCA) showed a clear separation between untreated and microbially treated pyrite along the PC1 dimension in both positive and negative ion spectra, with quintuplicate sites from the same sample clustering together (fig. S10). Further identification showed an increase in several biologically relevant organic molecular fragments after the biotic treatment [e.g., mass/charge ratio (m/z) 63 (CH[5]NO[2]^+), 149 (C[9]H[9]O[2]^+), and 72 (C[2]H[2]NO[2]^+)] ([112]Fig. 4 and table S1). The last one was previously identified on the surface of extracellular S^0 globules generated from sulfur oxidation by C. tepidum ([113]35). In contrast, peak intensities at m/z 56 (Fe^+) and 32 (S^−) were decreased ([114]Fig. 4). Fig. 3. SEM images and elemental maps. [115]Fig. 3. [116]Open in a new tab (A to C) SEM images indicative of a close physical association between C. tepidum and pyrite; (B) and (C) are enlargements of the yellow and blue rectangular areas in (A), respectively. (D to F) The elemental maps of C, Fe, and S for the area are shown in (A). Fig. 4. TOF-SIMS spectra. [117]Fig. 4. [118]Open in a new tab Chemical modification of pyrite surface after incubation with C. tepidum for 72 hours as indicated by the differences in normalized intensities (Δ normalized intensity) between treated and untreated pyrite samples as a function of m/z in both (A) positive and (B) negative ion spectra of TOF-SIMS. Positive “Δ normalized intensity” values indicate an increase in intensity after treatment with C. tepidum, while negative values indicate a decrease. X-ray photoelectron spectroscopy (XPS) analysis showed that iron on the surface of pyrite was still primarily in the form of ferrous iron [Fe(II)] after incubation with C. tepidum for 72 hours (i.e., reaction endpoint) (fig. S11), which aligned with previous findings that C. tepidum cannot use Fe(II) as an electron donor ([119]36). However, the valence state of sulfur on the surface of pyrite changed after incubation with C. tepidum for 24 hours, as indicated by the detection of SO[4]^2− ([120]Fig. 5). At this time point, sulfide had not been fully consumed ([121]Fig. 1B), and, thus, production of sulfate from oxidation of elemental sulfur or thiosulfate could not have occurred (fig. S3). Therefore, the presence of SO[4]^2− on the pyrite surface was attributed to the direct oxidation of sulfur within pyrite by C. tepidum ([122]26). Fig. 5. XPS spectra. [123]Fig. 5. [124]Open in a new tab The S 2p XPS spectra of (A) untreated and (B) treated (incubated with C. tepidum for 24 hours under the iron-depleted condition) pyrite. This time point (i.e., 24 hours) was selected to avoid the interference from the oxidized sulfur compounds generated from microbial oxidation of sulfide and thiosulfate. Transcriptomic response of C. tepidum to iron scarcity To reveal microbial coping strategies to iron scarcity, transcriptomic analysis was performed and compared across three treatments: iron-depleted treatment, aqueous Fe^2+ amendment (20 μM), and pyrite addition. The genome of C. tepidum is 2,154,946 base pairs in size, encoding about 2288 genes ([125]37). Results of transcriptomics showed that approximately 2103 genes were detected across the three treatments. PCA displayed a clear separation among the three treatments along the PC1 and PC2 dimensions, indicating treatment-specific gene expressions (fig. S12). Following normalization and statistical analysis using the Student’s t test, a comparative transcriptomic analysis was conducted with thresholds setting at log[2] fold change (log[2]FC) > 1 or log[2]FC < −1 to identify up-regulation and down-regulation of genes, respectively. P < 0.05 was taken as statistically significant. Overall, the addition of pyrite resulted in an up-regulation of 729 genes and a down-regulation of 446 genes compared with the iron-depleted treatment (fig. S13A). Furthermore, relative to the aqueous Fe^2+ treatment, the addition of pyrite resulted in an up-regulation of 549 genes and a down-regulation of 215 genes (fig. S13B). These results indicated that the presence of pyrite substantially affected the transcriptomic profile of C. tepidum. The Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment analysis revealed that sulfur metabolism pathways were substantially up-regulated in the presence of pyrite relative to the iron-depleted treatment ([126]Fig. 6A), consistent with the enhanced sulfur oxidation in this treatment ([127]Fig. 1B). In addition, pyrite addition resulted in the up-regulation of pathways related to vitamin metabolism, such as biosynthesis of folate (vitamin B9) and thiamine (B1) ([128]Fig. 6A). Folate is critical for DNA and RNA synthesis and cell proliferation ([129]38, [130]39), while thiamine diphosphate (i.e., biologically active form of thiamine) serves as a crucial cofactor in the tricarboxylic acid (Krebs) cycle ([131]40, [132]41). Furthermore, the pyrite treatment enriched the pathway involved in lipopolysaccharide biosynthesis ([133]Fig. 6A). Pathways related to the biosynthesis of amino acids (e.g., phenylalanine, tyrosine, tryptophan, arginine, and proline) were also enriched in the pyrite treatment when compared to either the iron-depleted ([134]Fig. 6A) or aqueous Fe^2+ treatment ([135]Fig. 6B). These molecular insights were well in line with the extensive detection of organics on pyrite surface ([136]Fig. 4). Fig. 6. KEGG pathway enrichment. [137]Fig. 6. [138]Open in a new tab Comparison of KEGG pathway enrichment of up-regulated gene clusters (A) pyrite treatment versus iron-depleted treatment and (B) pyrite treatment versus aqueous Fe^2+ treatment (20 μM). Dots represent pathway enrichment with color coding: Orange indicates high enrichment, and green indicates low enrichment. The size of the dots represents the gene number (count). Rich factor refers to the ratio of the up-regulated genes to all genes within a certain pathway. The larger value of a rich factor represents the higher level of enrichment. Compared to the treatment of aqueous Fe^2+, the riboflavin (B2) biosynthesis pathway was enriched in the pyrite treatment ([139]Fig. 6B). Riboflavin is a key component of the coenzymes flavin adenine dinucleotide and flavin mononucleotide, which is essential for energy production and electron transfer ([140]42–[141]44). In addition, two distinctive pathways related to nucleotide metabolism (i.e., pyrimidine metabolism and nucleotide excision repair) were also up-regulated in the pyrite treatment ([142]Fig. 6B). This phenomenon suggested an increased demand for DNA replication and repair, probably due to DNA damage caused by the cytotoxic reactive oxygen species produced from pyrite ([143]45, [144]46). From the view of gene expression, cytochrome-related genes such as CT1734 (cytochrome c), CT0073 (membrane-bound cytochrome c-[555]), cydA (cytochrome d ubiquinol oxidase), and CT0188 (c-type cytochrome) were among the most up-regulated genes in the pyrite treatment when compared with either the iron-depleted or aqueous Fe^2+ treatment (fig. S14). The up-regulation factor could reach up to 200-fold (log[2]FC = 7.66) in the pyrite treatment relative to the iron-depleted treatment (fig. S14). The substantial up-regulation of cytochrome-related genes implied that C. tepidum may secrete more cytochromes to enhance extracellular electron transfer (EET) ([145]47), possibly facilitating sulfur oxidation on the surface of pyrite under iron-depleted condition. In contrast, genes related to iron transport and iron storage were among the most down-regulated ones in the pyrite treatment relative to the iron-depleted treatment (fig. S14). For example, genes encoding ferrous iron transport protein such as CT1744 [ferrous iron transport protein A (feoA)], CT1741 (feoC), and feoB-1 were down-regulated up to 90-fold (log[2]FC = −6.52) in the pyrite treatment (fig. S14). Such substantial down-regulations likely reflected the rapid and efficient accumulation of cell-associated Fe in the presence of pyrite by the time of sampling for transcriptomic analysis (i.e., the log phase) ([146]Fig. 2, B and C), reducing the need for iron transport. In addition, the representative iron storage gene ftn, which encodes ferritin and is typically overexpressed in response to iron deficiency ([147]33), was also substantially down-regulated when pyrite was provided (fig. S14). This down-regulation also suggests that pyrite served as the primary and effective iron source, ensuring sufficient intracellular iron levels for microbial growth. Moreover, the relatively minor differences in the abundance of iron transport and storage genes between the aqueous Fe^2+ and pyrite treatments further support this possibility. Proteomic response of C. tepidum to iron scarcity Of a total of 2288 proteins predicted from the C. tepidum genome ([148]37), 371 proteins were shared across all three treatments (i.e., iron-depleted treatment, 20 μM aqueous Fe^2+ amendment, and pyrite addition). Thresholds of log[2]FC > 1 or log[2]FC < −1 and P < 0.05 were set to identify significant changes in protein expression (fig. S15). Results showed that relative to the iron-depleted treatment, 119 proteins were up-regulated and 39 proteins were down-regulated in the pyrite treatment (fig. S15A). When comparing pyrite treatment with aqueous Fe^2+ treatment, 57 proteins were up-regulated, and 19 proteins were down-regulated (fig. S15B). In particular, the expression of iron-related proteins (i.e., requiring iron as a cofactor or binding with iron) was analyzed because the primary aim of this study was to investigate whether C. tepidum can use iron within pyrite under iron-depleted conditions. Compared with the iron-depleted treatment, proteins involved in sulfur oxidation and photosynthesis, particularly those requiring iron or iron-sulfur clusters as cofactors, were up-regulated when pyrite was provided ([149]Fig. 7), consistent with the transcriptomic data ([150]Fig. 6) and the enhanced sulfur oxidation in this treatment ([151]Fig. 1B). For example, proteins such as aspB (adenylyl-sulfate reductase, beta subunit) and dsrK (DsrK protein) exhibited substantial up-regulations with log[2]FC > 2 ([152]Fig. 7). Notably, most of these proteins were also up-regulated when comparing pyrite treatment with aqueous Fe^2+ addition ([153]Fig. 7), which corresponded to the slightly faster cell growth ([154]Fig. 1A) and sulfur oxidation rate observed in the pyrite treatment ([155]Fig. 1B). In agreement with the transcriptomic results (fig. S14), iron transport (e.g., feoA-1 and feoB-1) and storage proteins (e.g., ferritin) were substantially down-regulated in the presence of pyrite compared with iron-depleted treatment ([156]Fig. 7). This likely reflected a dynamic response of cells to iron. Cells were initially forced to up-regulate iron transport and storage proteins under iron-depleted conditions to efficiently acquire iron from surrounding environments ([157]48, [158]49) but by the time of sampling for proteomic analysis (i.e., the log phase), intracellular iron may be sufficient to down-regulate these iron transport and storage proteins. In addition, the expression differences of proteins related to iron transport and storage were quite small between aqueous Fe^2+ and pyrite treatments ([159]Fig. 7), further suggesting that pyrite, similar to aqueous Fe^2+ with a high bioavailability, is an efficient iron source for C. tepidum in iron-depleted, euxinic environments. Fig. 7. Differential protein expressions. [160]Fig. 7. [161]Open in a new tab Differential expression of proteins related to sulfur oxidation, photosynthesis, and iron transport and storage in C. tepidum incubated under varying conditions. For proteins without official gene symbols, the corresponding gene ordered locus names are provided. Red color indicates log[2]FC > 0, while blue color indicates log[2]FC < 0. No pathway enrichment of down-regulated gene clusters was found to be significant (P < 0.05). DISCUSSION Importance of iron in the growth of anoxygenic phototrophs Given the pivotal roles of iron in photosynthetic reaction centers of oxygenic phototrophs, it has been well recognized that iron is one of the commonly limited nutrients for the growth of marine phytoplankton because aqueous iron tends to precipitate in oxic environments. Fluctuations of iron concentration largely control the magnitude and dynamics of oceanic primary production ([162]14, [163]15) and thus are of significance to both modern and paleoclimate changes because the metabolic activities of phytoplankton regulate the carbon exchange between atmosphere and ocean ([164]50, [165]51). However, before the advent of oxygenic photosynthesis, anoxygenic phototrophs such as GSB dominantly contribute to the primary production in oceans, especially during the Proterozoic ([166]17, [167]52). In general, anaerobes, especially these early-originated species, are not believed to suffer from iron deficiency because of the presumably high solubility and concentration of reduced iron in paleo-oceans. However, widespread and recurrent euxinia in Proterozoic oceans, where aqueous iron tends to precipitate with H[2]S as solid iron sulfide minerals ([168]2, [169]20), poses a great challenge of iron deficiency to these anoxygenic phototrophs thriving in sulfidic environments ([170]18, [171]20). Because of the importance of oxygenic phototrophs in oxygenating the early Earth and in controlling carbon sequestration in the modern ocean, extensive investigations about how oxygenic phototrophs cope with iron deficiency have been conducted. Nevertheless, little is known about how iron deficiency affects the growth of anoxygenic phototrophs. Such knowledge gap largely hinders our understanding of the biogeochemical cycles of carbon, sulfur, and nitrogen in sulfidic environments. Our results here demonstrate that iron deficiency suppressed the growth of C. tepidum, as evidenced by a reduction in biomass ([172]Fig. 1A and figs. S1 and S2) and impairment of sulfur oxidation ([173]Fig. 1B and fig. S3). Furthermore, cultivation experiments also suggest that iron concentrations above 0.2 μM are required to sustain active growth and sulfur oxidation of C. tepidum ([174]Fig. 1). However, such iron concentrations are higher than those typically found in natural euxinic environments where aqueous ferrous iron is scarce, yet GSB often thrive ([175]18, [176]20). This apparent paradox suggests that alternative iron sources such as mineral-associated iron may support the growth of anoxygenic phototrophs. In euxinic environments, reactive iron may be incorporated into fine-grained pyrite, creating a spatial copresence between the suspended pyrite and the niches of anoxygenic phototrophs ([177]1). Therefore, one possible explanation is that pyritic iron supports the growth and metabolism of GSB, considering the dominance of pyrite in the iron pool of euxinic settings. Our results demonstrated that C. tepidum can directly use pyritic iron to meet its cellular needs ([178]Fig. 2), enabling active growth and sulfur oxidation even when aqueous iron is scarce ([179]Fig. 1). However, note that a highly dynamic iron-sulfur cycle occurs in euxinic environments, particularly near or below the chemocline where other transient iron reservoirs (e.g., mackinawite, greigite, and FeS generated from the progressive sulfidation of iron oxides) may be present before complete pyritization ([180]20, [181]25). Given that GSB C. tepidum is capable of using iron in the pyrite (i.e., the thermodynamically stable phase of iron sulfide minerals), these metastable iron sulfide minerals possibly also serve as the key iron sources for the growth of anoxygenic phototrophs. Overall, our findings suggest that mineral-bound iron plays a previously underappreciated role in sustaining anoxygenic phototrophs in euxinic environments. Possible mechanisms of iron acquisition by GSB While iron is an irreplaceable element of many proteins and enzymes, previous studies have primarily focused on the iron acquisition strategies of aerobic microorganisms. In response to low iron availability, aerobic bacteria can secrete high-affinity extracellular ligands such as siderophores to bind and solubilize iron to meet their cellular iron demands ([182]49). In addition, aerobes can also uptake more iron than immediately needed, storing the excess for later use depending on the iron concentration in environments ([183]49). However, siderophores are generally not produced by anaerobes with a few exceptions ([184]53, [185]54). The genome of C. tepidum (i.e., the strictly anaerobic phototrophs applied in this study) does not contain any genes related to siderophore biosynthesis ([186]37). Consistently, no siderophores or siderophore-complexed iron was detected in any treatments, implying that C. tepidum adopts different strategies to meet its cellular iron requirements under aqueous iron-depleted conditions. Our findings suggest that C. tepidum acquired iron via oxidation of sulfur on pyrite surface. Specifically, C. tepidum modified the surface of pyrite with organic compounds such as lipopolysaccharides and amino acids, as supported by the presence of organic coatings on pyrite surface ([187]Fig. 4) and up-regulation of relevant pathways, genes, and proteins ([188]Figs. 6 and [189]7 and fig. S14). Such surface modifications may facilitate oxidation of pyritic sulfur through EET mediated by cytochromes ([190]Fig. 4 and fig. S14). It is known that C. tepidum can oxidize sulfide to S^0 globules and sulfate ([191]55, [192]56), consistent with our results (fig. S3). However, when solid-phase S^0 purchased from reagent companies was directly added into the medium, C. tepidum could not use these exogenous S^0 globules, despite their similar composition and structure to biogenic S^0 globules ([193]56). This biooxidizable property of biogenic S^0 globules has been attributed to their surface modifications by microbial metabolites ([194]35, [195]56) [e.g., molecular fragment with a m/z 72 (C[2]H[2]NO[2]^+)]. A similar organic coating was also observed on the pyrite surface as shown by the TOF-SIMS results in this study ([196]Fig. 4), suggesting that C. tepidum may modify the pyrite surface with comparable organics. This modification could facilitate the oxidation of sulfur within pyrite ([197]Fig. 5), thereby destabilizing the pyrite structure and promoting the release of ferrous iron into the surrounding environment. Furthermore, the requirement for direct contact between C. tepidum and pyrite surface ([198]Fig. 3 and fig. S8) aligns with the broader understanding that microorganisms often need to establish a physical connection with a mineral surface to exchange matter and energy ([199]21, [200]25, [201]47). Such mineral-microbe interaction may involve EET, where c-type cytochromes or other redox-active proteins mediate electron flow between the mineral and the microbial cell ([202]47). A recent study demonstrated that anoxygenic phototrophic PSB can oxidize sulfur within pyrite using certain monoheme c-type cytochromes as electron carriers under sulfur-limited conditions ([203]26). In addition, membrane-associated cytochromes play critical roles in the oxidization of extracellular solid-phase S^0 ([204]30). Therefore, together with the results that cytochrome-related genes were among the most up-regulated genes in the presence of pyrite (fig. S14), we speculate that C. tepidum can oxidize pyritic sulfur via EET through cytochromes, as indicated by the early and direct formation of sulfate on pyrite surface (i.e., not from oxidation of sulfide and thiosulfate; [205]Fig. 5), thereby destabilizing the pyrite structure and potentially facilitating its iron utilization. Although the exact mechanisms of how surface modification and EET facilitate microbial iron acquisition remain unclear, our findings offer insights into survival strategies of anaerobic microbes in iron-depleted environments. Implications for bioavailability of pyritic iron in anoxic environments Minerals and microbes have coevolved throughout much of Earth’s history ([206]32, [207]57). Various metabolic processes, including anoxygenic photosynthesis and sulfur oxidation, require enzymes that use specific transition metals such as iron for catalytic functions ([208]57, [209]58). Consequently, the origin and evolution of microbial metabolisms are tightly regulated by metal availability in the environment ([210]32, [211]57). The capability of microorganisms to use mineral-associated metals is a distinct ecological advantage that enables them to survive and function in diverse environments ([212]53, [213]59, [214]60). Our study emphasizes that pyrite can potentially alleviate the iron limitation faced by anaerobic microbial communities during oceanic euxinia in paleo-ocean. It has traditionally been believed that pyrite is stable under anoxic, low-temperature conditions, and chemical weathering of pyrite typically occurs in oxic environments, resulting in a large-scale input of sulfate to oceans ([215]22–[216]24). Nevertheless, a recent study showed that anoxic photochemical weathering of pyrite induced by sunlight exposure in the presence of aqueous Fe^2+ contributes substantial amounts of sulfate to the oceans during the anoxic late Archean ([217]61). In this reaction, aqueous Fe^2+ was photooxidized as Fe^3+, which can lead to the chemical oxidation of sulfur within pyrite, releasing Fe^2+ and sulfate into solutions ([218]61). Similarly, the traditional view that pyrite is only accessible to aerobic microbes or anaerobic microbes in the presence of oxidizing agents (e.g., Fe^3+, NO[3]^–, and MnO[2]) ([219]22–[220]24) is challenged by emerging evidences ([221]21, [222]25, [223]26). For example, anaerobic methanogenic archaea can use iron and sulfur from pyrite, possibly via reductive dissolution of pyrite ([224]21, [225]25). Under sulfur-limited conditions, anoxygenic phototroph PSB is capable of oxidizing pyritic sulfur, via EET, to polysulfide or elemental sulfur, but not to sulfate ([226]26). However, under the iron-limited condition applied in this study, pyrite surface–bound sulfur was likely oxidized to sulfate by GSB, facilitating its iron acquisition. In sulfidic habitats where anoxygenic phototrophs PSB and GSB flourish, their metabolic activities should be limited by iron, not sulfur. Therefore, our observed microbial oxidation of pyritic sulfur under iron-depleted conditions may be more environmentally relevant in both ancient and contemporary euxinia, enhancing the bioavailability of pyritic iron toward anoxygenic phototrophs. These results imply that microbial weathering of pyrite may be more prevalent on early Earth than previously thought and may have contributed to the biogeochemical cycles of elements such as carbon and sulfur in the paleo-ocean. MATERIALS AND METHODS Bacterial strain and cultivation medium The representative bacterial strain of GSB C. tepidum was obtained from the American Type Culture Collection (ATCC 49652) and used as a model anaerobic photosynthetic sulfur-oxidizing bacterium due to its well-characterized metabolic pathways and ecological relevance ([227]12, [228]13, [229]37). To investigate the role of iron in the growth and sulfur oxidation of C. tepidum, an iron-depleted medium was prepared, which consisted of the following (per liter): 0.5 g of KH[2]PO[4], 0.5 g of MgSO[4]·7H[2]O, 0.05 g of CaCl[2]·2H[2]O, 0.025 g of KCl, 1.0 g of NH[4]Cl, 2 ml of trace metal solution SL-6, 0.5 ml of Na-resazurin solution (0.1% w/v) as a redox indicator, and 100 mM 1,4-piperazinediethanesulfonic acid (PIPES) to buffer pH around 6.8. The trace metal solution SL-6 contained the following (per liter): 0.1 g of ZnSO[4]·7H[2]O, 0.03 g of MnCl[2]·4H[2]O, 0.3 g of H[3]BO[3], 0.2 g of CoCl[2]·6H[2]O, 0.01 g of CuCl[2]·2H[2]O, 0.02 g of NiCl[2]·6H[2]O, and 0.03 g of Na[2]WO[4]·2H[2]O. Although no Fe compounds were added to this iron-depleted medium, a trace amount of aqueous iron might present because of certain impurities in chemicals. In addition, a canonical medium with the addition of 5 ml of 0.1% ferric citrate solution (~20 μM Fe) was also prepared for initial bacterial growth and cell harvest. The prepared medium was aliquoted into 250-ml serum bottles (200 ml of solution and 50 ml of headspace) and purged with high-purity nitrogen (N[2], 99.999%) in a water bath (100°C) for 2.5 hours to remove dissolved oxygen. After autoclaving, the medium was amended with 1% (v/v) B[12] stock solution (0.0002%, filter sterilized and anoxic) in an anoxic glovebox (95% N[2] and 5% H[2], Coy Laboratory Products, Grass Lake, Michigan, USA). Before inoculation, sulfide (Na[2]S·9H[2]O) and thiosulfate (Na[2]S[2]O[3]·5H[2]O) were added from their anoxic, filter-sterilized stock solutions in an anoxic glovebox to achieve a final concentration of 2 mM each. In addition, NaHCO[3] (1.5 g/liter) was added as a carbon source for photoautotrophic growth of C. tepidum. Pretreatment of pyrite particles Pyrite (FeS[2]) was selected as a representative iron-containing mineral because of its abundance in ancient and contemporary euxinic environments, where anoxygenic phototrophic sulfur bacteria flourish ([230]20). A specimen pyrite (FeS[2]) sample, purchased from a mineral company (Pinquan, Zhejiang, China), was manually ground and size fractionated to obtain a 75- to 150-μm fraction. In addition, a smaller size fraction of 35.5 to 75 μm was separated to investigate the effect of particle size on cell growth and sulfur oxidation. Subsequently, the sieved pyrite particles were washed to remove any possible impurities according to the protocol described in a previous study ([231]21). Briefly, in an anoxic glovebox, the particles underwent four rinses with 1 N HCl, one rinse with 6 N HCl, three rinses with acetone, and, last, four rinses with Milli-Q water (18.2 megohms). The cleaned pyrite particles were freeze dried and preserved in an anoxic glovebox. Experimental setup All experiments were conducted in 50-ml serum bottles, which were presoaked in 5% nitric acid and rinsed three times with Milli-Q water to remove any impurities. In the following experiments, C. tepidum was inoculated into the serum bottles with a volume of 2%, and the experimental bottles were incubated at 47°C for 72 hours under white-light illumination (~2400 lux) ([232]12). Before the experiments with pyrite as the sole iron source, C. tepidum cells in the canonical medium were transferred three times into the iron-depleted medium to minimize any cellular carryover of iron. Subsequently, cells collected from the iron-depleted medium were inoculated again into the iron-depleted medium, followed by various treatments: (i) without the addition of iron; (ii) with the addition of various amounts of aqueous FeCl[2] stock solution (anoxic, filter sterilized) to achieve a final iron concentration of 0.2, 2, and 20 μM, respectively; and (iii) with the addition of pyrite particles (1 g/liter). These experiments were set up to investigate the importance of iron in cell growth and sulfur oxidation and to test whether solid-phase pyrite can serve as the sole iron source to support the metabolic activities of C. tepidum. In addition, three control experiments were designed: (i) quartz-substitution experiment (i.e., replacing pyrite with similar-sized, chemically inert quartz) to exclude the possibility that pyrite could enhance C. tepidum growth via the provision of attachment sites instead of serving as an iron source; (ii) pyrite-leachate experiment (i.e., pyrite was first soaked into the iron-depleted medium for 7 days and removed by filtering in an anoxic glovebox before inoculation of C. tepidum) to test whether iron from the abiotic dissolution of pyrite accounted for the growth promotion of C. tepidum; and (iii) dialysis bag experiment (i.e., pyrite was enclosed in a dialysis bag with a molecular weight cutoff of 500 Da) to examine the importance of direct contact between cell and pyrite to microbial growth and sulfur oxidation. OD, BChl c, and total protein concentration measurements At selected time points, subsamples were taken from the reaction bottles using a sterilized syringe inside an anoxic glovebox for the measurement of OD[750] to record cell growth ([233]13, [234]62). A previous study has demonstrated that the formation of extracellular elemental sulfur by C. tepidum did not interfere with OD measurement at this wavelength ([235]63). In addition, BChl c (a primary light-harvesting pigment produced by GSB) was extracted using methanol and quantified spectrophotometrically at 669 nm, applying an established extinction coefficient ([236]64). At the end of the reaction, the total cellular protein concentration was measured with the Bradford microassay to measure any biomass difference across various treatments ([237]65). Sulfide, thiosulfate, elemental sulfur, sulfite, and sulfate measurements At selected time points, subsamples were collected for concentration measurements of reactants (i.e., sulfide and thiosulfate) and products (i.e., elemental sulfur, sulfite, and sulfate) to monitor microbial sulfur oxidation. Specifically, sulfide concentration was measured with the methylene blue colorimetric assay ([238]66). Thiosulfate, sulfite, and sulfate concentrations were determined using ion chromatography equipped with an AS-14A anion-exchange column (4 mm by 250 mm; Thermo Fisher Scientific, Dionex IonPac), using 0.8 mM bicarbonate and 4.5 mM carbonate buffer as the eluent at a flow rate of 1 ml/min ([239]29). Extracellular elemental sulfur was firstly extracted with toluene ([240]67) and then measured with high-performance liquid chromatography (HPLC) equipped with a C[18] column (4.6 mm by 150 mm; Agilent, Eclipse Plus). The analysis was performed at a flow rate of 1 ml/min with a sample size of 20 μl, using a methanol/water mixture (98:2 v/v) as the eluent ([241]68, [242]69). Measurements of aqueous iron in culture supernatant and cell-associated iron To reflect cellular iron uptake during cell growth, concentrations of aqueous iron in culture supernatant and cell-associated iron were monitored with inductively coupled plasma mass spectrometry (ICP-MS; PerkinElmer, NexION 300X). Briefly, samples were taken from the reaction tubes using a sterilized syringe and centrifugated at a speed of 3500g for 10 min to separate supernatant from cell biomass. Iron concentration in the culture supernatant was measured directly, while the cell-associated iron was measured after digesting the collected cell pellets with the concentrated hydrochloric acid (i.e., 12 N HCl) for 15 min. In addition, Chrome Azurol S assay was conducted to check whether C. tepidum produced iron-binding ligands—siderophores ([243]60), and the presence or absence of siderophore-complexed iron was further discerned with LC-ICP-MS ([244]60). SEM observations To observe any morphological change of pyrite and its spatial relationship with C. tepidum cells, SEM observations were made for cell-pyrite suspensions. At the end of the reaction, suspensions of cell and pyrite were fixed for 90 min in a mixture of 2% paraformaldehyde and 2.5% glutaraldehyde buffered with phosphate buffer solution (130 mM NaCl and 10 mM sodium phosphate), followed by sequential dehydration using 25, 50, 75, 95, and 100% ethanol, successively ([245]70). Subsequently, samples were dried using a critical point dryer and coated with platinum (Pt) using a Quorum SC7620 sputter coater. SEM observations were conducted with a Zeiss SUPRA 55 SAPPHIRE scanning electron microscope equipped with a Genesis 2000 x-ray energy dispersive spectroscopy. The accelerating voltage was set to 10 to 15 kV, and the working distance was 15 cm. XRD analysis XRD was conducted to confirm the purity of the cleaned pyrite and to check whether pyrite was altered or any secondary minerals were formed at the end of reaction. Analysis was performed with a Rigaku SmartLab x-ray powder diffractometer using a rotating anode generator (Cu Kα radiation) with a power of 9000 W (200 kV, 45 mA). Samples were scanned over a 2θ range of 3° to 90°, stepping at 0.02° with a counting time of 1 s per step. MDI Jade 6 software was used to identify mineral phases. TOF-SIMS analysis To reveal any compositional changes on the surface of pyrite at the end of the reaction, static TOF-SIMS, a surface-sensitive technique with a penetration depth of 1 to 2 nm ([246]71), was applied with a TOF.SIMS5 spectrometer (IONTOF GmbH, Münster, Germany). A pulsed Bi[3]^+ primary ion beam of 25 keV was used to collect the TOF-SIMS spectra. Five positive ion spectra and five negative ion spectra were collected per sample from randomly selected surface sites, with minimum counts of 1000 and a minimum signal-to-noise ratio of 3. After background correction and normalization of obtained mass spectrum data, PCA was performed to evaluate any differences in molecular clusters between the untreated and microbially treated pyrite. XPS analysis XPS analysis, a surface-sensitive technique with a penetration depth up to 10 nm ([247]72), was applied to determine the valence states of iron and sulfur on the surface of microbially reacted pyrite. The XPS data were collected using a Thermo Fisher Scientific ESCALAB 250Xi spectrometer with a monochromatic Al Kα x-ray source. The samples were prepared within an anoxic glovebox to avoid any potential air oxidation of pyrite. High-resolution spectra were recorded in 0.1-eV steps for iron and 0.05-eV steps for sulfur with a pass energy of 35 eV. The spectra were analyzed with Thermo Fisher Scientific Avantage software. The collected binding energy values were calibrated using carbon (C 1s) as the reference (i.e., 284.60 eV) ([248]73). The Fe 2p[3/2] and Fe 2p[1/2] peaks were fitted with binding energies of 707.15 and 720.05 eV, respectively ([249]73). The S 2p spectra were fitted using spin-orbit split doublets, with a separation of 1.18 eV between the peaks ([250]74). Both components were constrained to have the same full width at half maximum, and the area of the S 2p[3/2] peak was set to be twice that of the S 2p[1/2] peak ([251]26, [252]74). The S 2p[3/2] binding energies for monosulfide (S^2−), disulfide (S[2]^2−), polysulfide (S[n]^2−), and sulfate (SO[4]^2−) were assigned to 161.14, 162.10, 164.10, and 168.10 eV, respectively ([253]26, [254]75). Accordingly, the S 2p[1/2] binding energies for S^2−, S[2]^2−, S[n]^2−, and SO[4]^2− were assigned to 162.32, 163.28, 165.28, and 169.28 eV, respectively ([255]26, [256]75). Transcriptomic and proteomic analyses Cells at the logarithmic phase (30 hours) were harvested for transcriptomic and proteomic analyses. These analyses were performed and compared across three treatments: the iron-depleted treatment, the 20 μM aqueous Fe^2+ treatment, and the pyrite treatment. For transcriptomic analysis, total RNA was extracted from samples using the TRIzol method ([257]76). RNA quality was assessed with a Thermo Fisher Scientific NanoDrop One equipped with an Agilent 4200 tape station. Ribosomal RNA (rRNA) was removed with the Epicentre Ribo-Zero rRNA Removal Kit. Libraries were constructed with the NEBNext Ultra II Directional RNA Library Prep Kit for Illumina, followed by quality control and sequencing on the Illumina NovaSeq platform (PE150). Raw sequencing reads were obtained in FASTQ format and subjected to quality control using fastp ([258]77). Clean reads were aligned to the C. tepidum genome using Bowtie 2, and transcript abundance was quantified ([259]78). PCA was performed to show the overall expression patterns of samples across different treatments. Differential expression analysis was performed using DESeq2 ([260]79). An R package, clusterProfiler, was used to conduct enrichment analysis for the KEGG pathway ([261]80). For proteomic analysis, total protein was extracted from cells by sonication, followed by separation using SDS–polyacrylamide gel electrophoresis ([262]62). The protein bands were excised from the gel, digested with trypsin, and analyzed by electrospray ionization Orbitrap MS. Peptide sequences were matched against the UniProt protein database derived from the C. tepidum genome using MaxQuant software (version 2.2.6) to identify and quantify proteins. Filtering of potential contaminants, z-score normalization, Student’s t test, and FC analysis were performed in Perseus (version 2.2.1). Acknowledgments