Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Distinct gene clusters drive formation of ferrosome organelles in bacteria

Abstract

Cellular iron homeostasis is vital and maintained through tight regulation of iron import, efflux, storage and detoxification1,2,3. The most common modes of iron storage use proteinaceous compartments, such as ferritins and related proteins4,5. Although lipid-bounded iron compartments have also been described, the basis for their formation and function remains unknown6,7. Here we focus on one such compartment, herein named the ‘ferrosome’, that was previously observed in the anaerobic bacterium Desulfovibrio magneticus6. Using a proteomic approach, we identify three ferrosome-associated (Fez) proteins that are responsible for forming ferrosomes in D. magneticus. Fez proteins are encoded in a putative operon and include FezB, a P1B-6-ATPase found in phylogenetically and metabolically diverse species of bacteria and archaea. We show that two other bacterial species, Rhodopseudomonas palustris and Shewanella putrefaciens, make ferrosomes through the action of their six-gene fez operon. Additionally, we find that fez operons are sufficient for ferrosome formation in foreign hosts. Using S. putrefaciens as a model, we show that ferrosomes probably have a role in the anaerobic adaptation to iron starvation. Overall, this work establishes ferrosomes as a new class of iron storage organelles and sets the stage for studying their formation and structure in diverse microorganisms.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Proteins enriched with ferrosomes are essential for ferrosome formation.
Fig. 2: fez genes are essential for ferrosome formation and function in S. putrefaciens.
Fig. 3: fez genes enable ferrosome formation in foreign hosts.

Similar content being viewed by others

Data availability

The mass spectrometry proteomics data have been deposited to the ProteomeXchange Consortium via the PRIDE67 partner repository with dataset identifier PXD017470. Ferrosome-associated proteins presented in Fig. 1a were identified from the data in Supplementary Table 11. The sequences, alignment and tree data used to generate Fig. 1g are provided as Supplementary Data 2. KEGG60 and IMG/M ER52 were used to collect data.

References

  1. Andrews, S. C. In Advances in Microbial Physiology vol. 40 (ed. Poole, R. K.) 281–351 (Academic, 1998).

  2. Touati, D. Iron and oxidative stress in bacteria. Arch. Biochem. Biophys. 373, 1–6 (2000).

    Article  CAS  PubMed  Google Scholar 

  3. Andrews, S. C., Robinson, A. K. & Rodríguez-Quiñones, F. Bacterial iron homeostasis. FEMS Microbiol. Rev. 27, 215–237 (2003).

    Article  CAS  PubMed  Google Scholar 

  4. Andrews, S. C. The ferritin-like superfamily: evolution of the biological iron storeman from a rubrerythrin-like ancestor. Biochim. Biophys. Acta 1800, 691–705 (2010).

    Article  CAS  PubMed  Google Scholar 

  5. Nichols, R. J., Cassidy-Amstutz, C., Chaijarasphong, T. & Savage, D. F. Encapsulins: molecular biology of the shell. Crit. Rev. Biochem. Mol. Biol. 52, 583–594 (2017).

    Article  CAS  PubMed  Google Scholar 

  6. Byrne, M. E. et al. Desulfovibrio magneticus RS-1 contains an iron- and phosphorus-rich organelle distinct from its bullet-shaped magnetosomes. Proc. Natl Acad. Sci. USA 107, 12263–12268 (2010).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  7. Glasauer, S., Langley, S. & Beveridge, T. J. Intracellular iron minerals in a dissimilatory iron-reducing bacterium. Science 295, 117–119 (2002).

    Article  CAS  PubMed  ADS  Google Scholar 

  8. Sakaguchi, T., Arakaki, A. & Matsunaga, T. Desulfovibrio magneticus sp. nov., a novel sulfate-reducing bacterium that produces intracellular single-domain-sized magnetite particles. Int. J. Syst. Evol. Microbiol. 52, 215–221 (2002).

    Article  CAS  PubMed  Google Scholar 

  9. Grant, C. R., Rahn-Lee, L., LeGault, K. N. & Komeili, A. Genome editing method for the anaerobic magnetotactic bacterium Desulfovibrio magneticus RS-1. Appl. Env. Microbiol. 84, e01724-18 (2018).

    Article  ADS  Google Scholar 

  10. Rahn-Lee, L. et al. A genetic strategy for probing the functional diversity of magnetosome formation. PLoS Genet. 11, e1004811 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  11. Argüello, J. M., Eren, E. & González-Guerrero, M. The structure and function of heavy metal transport P1B-ATPases. BioMetals 20, 233 (2007).

    Article  PubMed  CAS  Google Scholar 

  12. Smith, A. T., Smith, K. P. & Rosenzweig, A. C. Diversity of the metal-transporting P1B-type ATPases. J. Biol. Inorg. Chem. 19, 947–960 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Chan, H. et al. The P-type ATPase superfamily. J. Mol. Microbiol. Biotechnol. 19, 5–104 (2010).

    CAS  PubMed  Google Scholar 

  14. Padilla-Benavides, T., Long, J. E., Raimunda, D., Sassetti, C. M. & Argüello, J. M. A novel P1B-type Mn2+-transporting ATPase is required for secreted protein metallation in mycobacteria. J. Biol. Chem. 288, 11334–11347 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Botella, H. et al. Mycobacterial P1-type ATPases mediate resistance to zinc poisoning in human macrophages. Cell Host Microbe 10, 248–259 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Russ, W. P. & Engelman, D. M. The GxxxG motif: a framework for transmembrane helix–helix association. J. Mol. Biol. 296, 911–919 (2000).

    Article  CAS  PubMed  Google Scholar 

  17. Unterreitmeier, S. et al. Phenylalanine promotes interaction of transmembrane domains via GxxxG motifs. J. Mol. Biol. 374, 705–718 (2007).

    Article  CAS  PubMed  Google Scholar 

  18. Jarsch, I. K., Daste, F. & Gallop, J. L. Membrane curvature in cell biology: an integration of molecular mechanisms. J. Cell Biol. 214, 375–387 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Glasauer, S. et al. Mixed-valence cytoplasmic iron granules are linked to anaerobic respiration. Appl. Environ. Microbiol. 73, 993–996 (2007).

    Article  CAS  PubMed  ADS  Google Scholar 

  20. VerBerkmoes, N. C. et al. Determination and comparison of the baseline proteomes of the versatile microbe Rhodopseudomonas palustris under its major metabolic states. J. Proteome Res. 5, 287–298 (2006).

    Article  CAS  PubMed  Google Scholar 

  21. Rey, F. E. & Harwood, C. S. FixK, a global regulator of microaerobic growth, controls photosynthesis in Rhodopseudomonas palustris. Mol. Microbiol. 75, 1007–1020 (2010).

    Article  CAS  PubMed  Google Scholar 

  22. Bose, A. & Newman, D. K. Regulation of the phototrophic iron oxidation (pio) genes in Rhodopseudomonas palustris TIE-1 is mediated by the global regulator, FixK. Mol. Microbiol. 79, 63–75 (2011).

    Article  CAS  PubMed  Google Scholar 

  23. Amor, M. et al. Magnetotactic bacteria accumulate a large pool of iron distinct from their magnetite crystals. Appl. Environ. Microbiol. 86, e01278-20 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  24. Abdul-Tehrani, H. et al. Ferritin mutants of Escherichia coli are iron deficient and growth impaired, and fur mutants are iron deficient. J. Bacteriol. 181, 1415–1428 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Rolfe, M. D. et al. Lag phase is a distinct growth phase that prepares bacteria for exponential growth and involves transient metal accumulation. J. Bacteriol. 194, 686–701 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Bender, K. S. et al. Analysis of a ferric uptake regulator (Fur) mutant of Desulfovibrio vulgaris Hildenborough. Appl. Environ. Microbiol. 73, 5389–5400 (2007).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  27. Uebe, R. et al. Deletion of a fur-like gene affects iron homeostasis and magnetosome formation in Magnetospirillum gryphiswaldense. J. Bacteriol. 192, 4192–4204 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Wang, Q. et al. Physiological characteristics of Magnetospirillum gryphiswaldense MSR-1 that control cell growth under high-iron and low-oxygen conditions. Sci. Rep. 7, 2800 (2017).

    Article  PubMed  PubMed Central  ADS  CAS  Google Scholar 

  29. Pereira, P. M. et al. Transcriptional response of Desulfovibrio vulgaris Hildenborough to oxidative stress mimicking environmental conditions. Arch. Microbiol. 189, 451–461 (2008).

    Article  CAS  PubMed  Google Scholar 

  30. Zhou, A. et al. Hydrogen peroxide-induced oxidative stress responses in Desulfovibrio vulgaris Hildenborough. Environ. Microbiol. 12, 2645–2657 (2010).

    CAS  PubMed  Google Scholar 

  31. Caffrey, S. M. & Voordouw, G. Effect of sulfide on growth physiology and gene expression of Desulfovibrio vulgaris Hildenborough. Antonie Van Leeuwenhoek 97, 11–20 (2010).

    Article  CAS  PubMed  Google Scholar 

  32. Ho, T. D. & Ellermeier, C. D. Ferric uptake regulator Fur control of putative iron acquisition systems in Clostridium difficile. J. Bacteriol. 197, 2930–2940 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Jiao, Y. & Newman, D. K. The pio operon is essential for phototrophic Fe(II) oxidation in Rhodopseudomonas palustris TIE-1. J. Bacteriol. 189, 1765–1773 (2007).

    Article  CAS  PubMed  Google Scholar 

  34. Kim, M.-K. & Harwood, C. S. Regulation of benzoate-CoA ligase in Rhodopseudomonas palustris. FEMS Microbiol. Lett. 83, 199–203 (1991).

    Article  CAS  Google Scholar 

  35. Guan, G. et al. PfeT, a P1B4-type ATPase, effluxes ferrous iron and protects Bacillus subtilis against iron intoxication. Mol. Microbiol. 98, 787–803 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Komeili, A., Vali, H., Beveridge, T. J. & Newman, D. K. Magnetosome vesicles are present before magnetite formation, and MamA is required for their activation. Proc. Natl Acad. Sci. USA 101, 3839–3844 (2004).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  37. Murat, D., Quinlan, A., Vali, H. & Komeili, A. Comprehensive genetic dissection of the magnetosome gene island reveals the step-wise assembly of a prokaryotic organelle. Proc. Natl Acad. Sci. USA 107, 5593–5598 (2010).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  38. Rey, F. E., Oda, Y. & Harwood, C. S. Regulation of uptake hydrogenase and effects of hydrogen utilization on gene expression in Rhodopseudomonas palustris. J. Bacteriol. 188, 6143–6152 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Clark, I. C. et al. Synthetic and evolutionary construction of a chlorate-reducing Shewanella oneidensis MR-1. mBio 6, e00282-15 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  40. Plumb, R. S. et al. UPLC/MSE: a new approach for generating molecular fragment information for biomarker structure elucidation. Rapid Commun. Mass Spectrom. 20, 1989–1994 (2006).

    Article  CAS  PubMed  ADS  Google Scholar 

  41. Geromanos, S. J., Hughes, C., Ciavarini, S., Vissers, J. P. C. & Langridge, J. I. Using ion purity scores for enhancing quantitative accuracy and precision in complex proteomics samples. Anal. Bioanal. Chem. 404, 1127–1139 (2012).

    Article  CAS  PubMed  Google Scholar 

  42. Shliaha, P. V., Bond, N. J., Gatto, L. & Lilley, K. S. Effects of traveling wave ion mobility separation on data independent acquisition in proteomics studies. J. Proteome Res. 12, 2323–2339 (2013).

    Article  CAS  PubMed  Google Scholar 

  43. Levin, Y. & Bahn, S. LC–MS/MS in proteomics. Methods Mol. Biol 658, 217–231 (2010).

    Article  CAS  PubMed  Google Scholar 

  44. Neilson, K. A. et al. Less label, more free: approaches in label-free quantitative mass spectrometry. Proteomics 11, 535–553 (2011).

    Article  CAS  PubMed  Google Scholar 

  45. Nahnsen, S., Bielow, C., Reinert, K. & Kohlbacher, O. Tools for label-free peptide quantification. Mol. Cell. Proteomics 12, 549–556 (2013).

    Article  CAS  PubMed  Google Scholar 

  46. Light, S. H. et al. A flavin-based extracellular electron transfer mechanism in diverse Gram-positive bacteria. Nature 562, 140–144 (2018).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  47. Engström, P. et al. Evasion of autophagy mediated by Rickettsia surface protein OmpB is critical for virulence. Nat. Microbiol. 4, 2538–2551 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Pinals, R. L. et al. Quantitative protein corona composition and dynamics on carbon nanotubes in biological environments. Angew. Chem. Int. Ed. 59, 23668–23677 (2020).

    Article  CAS  Google Scholar 

  49. Engström, P., Burke, T. P., Tran, C. J., Iavarone, A. T. & Welch, M. D. Lysine methylation shields an intracellular pathogen from ubiquitylation and autophagy. Sci. Adv. 7, eabg2517 (2021).

    Article  PubMed  PubMed Central  ADS  CAS  Google Scholar 

  50. Schneider, C. A., Rasband, W. S. & Eliceiri, K. W. NIH Image to ImageJ: 25 years of image analysis. Nat. Methods 9, 671–675 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Atolia, E. et al. Environmental and physiological factors affecting high-throughput measurements of bacterial growth. mBio 11, e01378-20 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  52. Chen, I.-M. A. et al. IMG/M v.5.0: an integrated data management and comparative analysis system for microbial genomes and microbiomes. Nucleic Acids Res. 47, D666–D677 (2018).

    Article  PubMed Central  CAS  Google Scholar 

  53. Kumar, S., Nei, M., Dudley, J. & Tamura, K. MEGA: a biologist-centric software for evolutionary analysis of DNA and protein sequences. Brief. Bioinform. 9, 299–306 (2008).

    Article  CAS  PubMed  Google Scholar 

  54. Castresana, J. Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Mol. Biol. Evol. 17, 540–552 (2000).

    Article  CAS  PubMed  Google Scholar 

  55. Stamatakis, A. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30, 1312–1313 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Lefort, V., Longueville, J.-E. & Gascuel, O. SMS: smart model selection in PhyML. Mol. Biol. Evol. 34, 2422–2424 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Letunic, I. & Bork, P. Interactive Tree Of Life (iTOL) v4: recent updates and new developments. Nucleic Acids Res. 47, W256–W259 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Steinegger, M. & Söding, J. MMseqs2 enables sensitive protein sequence searching for the analysis of massive data sets. Nat. Biotechnol. 35, 1026–1028 (2017).

    Article  CAS  PubMed  Google Scholar 

  59. Markiel, S. P. et al. Cytoscape: a software environment for integrated models of biomolecular interaction networks. Genome Res. 13, 2498–2504 (2003).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  60. Kanehisa, M. & Goto, S. Kyoto Encyclopedia of Genes and Genomes. Nucleic Acids Res. 28, 27–30 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Morris, J. H. et al. clusterMaker: a multi-algorithm clustering plugin for Cytoscape. BMC Bioinf. 12, 436 (2011).

    Article  CAS  Google Scholar 

  62. Sievers, F. et al. Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol. Syst. Biol. 7, 539 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  63. Finn, R. D., Clements, J. & Eddy, S. R. HMMER web server: interactive sequence similarity searching. Nucleic Acids Res. 39, W29–W37 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Eddy, S. R. Accelerated profile HMM searches. PLoS Comput. Biol. 7, e1002195 (2011).

    Article  MathSciNet  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  65. Bernsel, A., Viklund, H., Hennerdal, A. & Elofsson, A. TOPCONS: consensus prediction of membrane protein topology. Nucleic Acids Res. 37, W465–W468 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Crooks, G. E. WebLogo: a sequence logo generator. Genome Res. 14, 1188–1190 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Perez-Riverol, Y. et al. The PRIDE database and related tools and resources in 2019: improving support for quantification data. Nucleic Acids Res. 47, D442–D450 (2019).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank faculty at the EM-Lab at the University of California, Berkeley, for their assistance with TEM; A. Campos for assistance with the EDS measurements;  M. Pang for assistance in culturing R. palustris for ferrosome production; K. LeGault and H. McCausland for their help with conjugations in R. palustris and M. magneticum; J. Gralnick (University of Minnesota) for providing S. putrefaciens strain CN-32; and the Coates laboratory and the Niyogi laboratory for sharing equipment. The QB3/Chemistry Mass Spectrometry Facility at the University of California, Berkeley, received support from the National Institutes of Health (shared instrumentation grant 1S10OD020062-01). Research reported in this publication was supported by funding from the National Institutes of Health (R01GM084122 and R35GM127114), the Office of Naval Research (N000141310421) and the Bakar Fellows Program. H.A.T. is supported by the National Science Foundation Graduate Research Fellowship Program under grant DGE 1752814. M.A. is supported by a grant through the Fondation pour la Recherche Médicale (ARF201909009123).

Author information

Authors and Affiliations

Authors

Contributions

C.R.G. and A.K. conceived and designed the study. C.R.G. performed all molecular cloning, genetic manipulation, TEM, cellular fractionations and sample preparations for LC–MS analyses. A.T.I. performed all LC–MS analyses. C.R.G. identified ferrosome-associated proteins with assistance from A.T.I. and H.A.T. C.R.G. carried out the bioinformatic analyses and tree construction. M.A. performed the EDS experiments and analysis. C.R.G. performed all growth assays with assistance from S.K. C.R.G. and A.K. prepared the manuscript with input from S.K., A.T.I., H.A.T. and M.A.

Corresponding author

Correspondence to Arash Komeili.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature thanks the anonymous reviewers for their contribution to the peer review of this work.

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 Ferrosomes are visible by TEM in whole D. magneticus cells after transitioning from iron limited to iron replete conditions.

D. magneticus cells initially grown without iron (a) are shown 0.5 (b), 1.5 (c), and 6 (d) hours after addition of 100 μM ferric malate. (e) The maximum diameter of ferrosomes represented in b-d. Each data point represents one ferrosome and the bar indicates the mean maximum diameter in nm. Micrographs of D. magneticus one hour after adding low to high concentrations of ferric malate—1 μM (f), 10 μM (g), 100 μM (h), and 1 mM (i)—to iron-starved cells. (j) The maximum diameter of ferrosomes represented in f-i. Each data point represents one ferrosome and the bar indicates the mean maximum diameter in nm. Scale bars, 200 nm.

Extended Data Fig. 2 Isolation of ferrosomes and characteristics of associated proteins.

(a) Ferrosomes from ∆MAI D. magneticus (left) and magnetosomes from WT D. magneticus (right) form a pellet through 65% sucrose. Transmission electron micrographs of the ferrosome pellet (b) and the magnetosome pellet (c). Scale bars, 100 nm. (d-f) Membrane domain predictions of ferrosome-associated proteins in D. magneticus. DMR_28320 (a), DMR_28330 (b), and DMR_28340 (c) have 1, 5–6, and 0–2 putative transmembrane domains, respectively, as predicted by various methods analyzed through TOPCONS 1.065.

Extended Data Fig. 3 Multiple sequence alignment of FezB with characterized P1B-ATPases.

Conserved functional motifs in the actuator domain and the ATP-binding domain are indicated with blue and purple stars, respectively. The CxxC and histidine-rich metal binding sites in the cytoplasmic N-terminal domain of ZntA, CopA, and CopB are boxed. Transmembrane regions, predicted using TOPCONS 1.065, are underlined for each sequence. Putative metal-binding sites in the transmembrane domains are indicated with black stars.

Extended Data Fig. 4 WT and ∆fezBC D. magneticus strains make ferrosomes in iron replete medium when expressing fezABC in trans.

Transmission electron micrographs of WT (a) and ∆fezBC (b) strains with a control plasmid make magnetosomes (white carets) when grown in iron replete medium. When expressing fezABC in trans, both the WT (c) and ∆fezBC (d) strains make magnetosomes as well as ferrosomes when grown in iron replete medium. Areas of the cell containing one or more putative ferrosomes are indicated with yellow circles. Scale bars, 200 nm.

Extended Data Fig. 5 Sequence similarity network of proteins encoded by fez gene clusters and genes frequently found near fez gene clusters.

(a) Conserved fez gene clusters that encode FezB homologs. Conserved genes within the clusters are colored black. Gene clusters were identified using the “Gene cluster” tool in KEGG for each FezB homolog, in bold: Dde_0495, Dde_0498, Thimo_2900, vfu_A02104, SMUL_2748, RPA2333, KN400_3199, DMR_28330, and EUBELI_00578. The second copy of FezB in D. alaskensis, Dde_0498, is not shown because it is not part of a predicted conserved gene cluster. (b, c) Sequence similarity network highlighting the proteins encoded by ten genes upstream and downstream of 304 FezB homologs. Each node represents a protein and edges represent protein similarities that meet the specified e-value cutoff. (b) Network containing fez gene cluster-encoded proteins. Each group (labeled 1–8) contains one or more proteins encoded by conserved genes identified in (a) which are represented by black nodes and are labeled. Proteins or domains with an annotated function are labeled. Groups of proteins were further divided into subgroups which were used to identify proteins with GxxxG motifs in groups 2 and 5 and proteins with R-rich motifs in groups 1 and 3 (see Methods). The proteins represented in this network and their group/subgroup are listed in Supplementary Tables 36. (c) Network of proteins encoded by genes that are frequently found upstream and downstream of fez gene clusters. Only groups of more than 30 proteins are shown and the protein or domain annotation is labeled. Proteins with a known role in iron homeostasis are common and include iron transporters (FeoA, FeoB, outermembrane siderophore receptors, and some ABC transporters) and regulators (Fur and DtxR). The proteins represented in this network are listed in Supplementary Table 7.

Extended Data Fig. 6 Consensus motifs and characteristics of proteins with R-rich and GxxxG motifs.

Representative proteins encoded by fez gene clusters with (a) an R-rich motif or (b) a GxxxG motif. Logo shows the consensus motif for the subgroup or group of proteins to which the representative protein belongs. Predicted protein structure schematics show approximate location of the R-rich motif, putative transmembrane helices, and GxxxG motif for each protein (not to scale).

Extended Data Fig. 7 Transmission electron micrographs of S. putrefaciens and R. palustris.

WT S. putrefaciens (a, b) and R. palustris (c, d), ∆fezSp (e, f), ∆fezRp (g), ∆fezSp::fezSp (h, i), and ∆fezRp::fezRp (j). S. putrefaciens strains respiring fumarate in medium supplemented with 100 μM ferric malate (a, e, i) or 1 mM ferrous iron (b, f, j). R. palustris strains grown anaerobically (d, g, h) or aerobically (c). White arrows denote ferrosomes. Polyphosphate granules are indicated with white asterisks. Scale bars, 200 nm.

Extended Data Fig. 8 EDS spectra of S. putrefaciens and E. coli.

EDS spectrum of an S. putrefaciensfezSp cell, which does not form ferrosomes. (b, c) EDS spectra of S. putrefaciens WT obtained from an area in the cell that contained ferrosomes (b) and an area that had no visible ferrosomes (c). The red asterisk indicates the iron peak associated with ferrosomes in WT S. putrefaciens. (d, e) Spectra of the background taken from areas of the S. putrefaciens WT (d) and ∆fezSp (e) grids that contained no cells. (f) An EDS spectrum of E. coli fezSp+ obtained from an area in the cell that had no visible ferrosomes. (g, h) Spectra of the background taken from areas of the E. coli cells with a control plasmid (g) or E. coli fezSp+ (h).

Extended Data Fig. 9 Effect of EDTA on the growth of S. putrefaciens.

(a) OD595 measurements over time of S. putrefaciens WT (navy) and ∆fezSp (yellow) grown aerobically with the indicated concentrations of EDTA. Each line is the mean of 3 individual cultures (technical replicates); error bars indicate s.d. (b) OD595 measurements over time of S. putrefaciens WT (navy) and ∆fezSp (yellow) grown anaerobically with the indicated concentrations of EDTA. Each line is the mean of 6 individual cultures (2 biological replicates with 3 technical replicates, with the exception of 150 μM EDTA which had 2 technical replicates); error bars indicate s.d. (c) Growth rate versus OD595 of the individual cultures shown in (b). Each circle represents the growth rate for an individual culture.

Supplementary information

Reporting Summary

Supplementary Table 1

P-type ATPases and their characteristics used to generate Fig. 2a.

Supplementary Table 2

Characteristics of bacteria and archaea with a FezB homologue.

Supplementary Table 3

Proteins belonging to group 1 (P-type ATPases).

Supplementary Table 4

Proteins belonging to group 2 (GXXXG motif).

Supplementary Table 5

Proteins belonging to group 3 (R-rich motif/DUF4405).

Supplementary Table 6

Other conserved proteins belonging to groups 4–8.

Supplementary Table 7

Proteins associated with fez gene clusters.

Supplementary Table 8

Bacterial strains used in this study.

Supplementary Table 9

Plasmids used in this study.

Supplementary Table 10

List of primers used in this study.

Supplementary Table 11

Relative protein quantification of proteins enriched with ferrosomes.

Supplementary Data 1

Raw data for S. putrefaciens growth assays used for Fig. 2e–g and Extended Data Fig. 9.

Supplementary Data 2

Sequences, alignment and tree data used to generate Fig. 1g.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Grant, C.R., Amor, M., Trujillo, H.A. et al. Distinct gene clusters drive formation of ferrosome organelles in bacteria. Nature 606, 160–164 (2022). https://doi.org/10.1038/s41586-022-04741-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-022-04741-x

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing Microbiology

Sign up for the Nature Briefing: Microbiology newsletter — what matters in microbiology research, free to your inbox weekly.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing: Microbiology