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:

An allosteric redox switch involved in oxygen protection in a CO2 reductase

Abstract

Metal-dependent formate dehydrogenases reduce CO2 with high efficiency and selectivity, but are usually very oxygen sensitive. An exception is Desulfovibrio vulgaris W/Sec-FdhAB, which can be handled aerobically, but the basis for this oxygen tolerance was unknown. Here we show that FdhAB activity is controlled by a redox switch based on an allosteric disulfide bond. When this bond is closed, the enzyme is in an oxygen-tolerant resting state presenting almost no catalytic activity and very low formate affinity. Opening this bond triggers large conformational changes that propagate to the active site, resulting in high activity and high formate affinity, but also higher oxygen sensitivity. We present the structure of activated FdhAB and show that activity loss is associated with partial loss of the metal sulfido ligand. The redox switch mechanism is reversible in vivo and prevents enzyme reduction by physiological formate levels, conferring a fitness advantage during O2 exposure.

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: Enzymatic activity of FdhAB and variants.
Fig. 2: EPR spectra of WV species in WT FdhAB enzyme, C872A and M405A variants.
Fig. 3: Structural changes induced by the allosteric cleavage of the C845–C872 bond.
Fig. 4: Effect of oxygen on catalysis by WT FdhAB and C872A variant.
Fig. 5: Reversibility of the FdhAB redox switch.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available within the main text and its Supplementary Information file. The atomic coordinates and structure factors for the D. vulgaris H C872A variant structures have been deposited in the PDB under accession codes 8CM4, 8CM5, 8CM6 and 8CM7. Source data are provided with this paper.

References

  1. Dalle, K. E. et al. Electro- and solar-driven fuel synthesis with first row transition metal complexes. Chem. Rev. 119, 2752–2875 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Wang, G. et al. Electrocatalysis for CO2 conversion: from fundamentals to value-added products. Chem. Soc. Rev. 50, 4993–5061 (2021).

    Article  CAS  PubMed  Google Scholar 

  3. Zhang, S., Fan, Q., Xia, R. & Meyer, T. J. CO2 reduction: from homogeneous to heterogeneous electrocatalysis. Acc. Chem. Res. 53, 2–11 (2019).

    Google Scholar 

  4. Shi, J. et al. Enzymatic conversion of carbon dioxide. Chem. Soc. Rev. 44, 5981–6000 (2015).

    Article  CAS  PubMed  Google Scholar 

  5. Yishai, O., Lindner, S. N., Gonzalez de la Cruz, J., Tenenboim, H. & Bar-Even, A. The formate bio-economy. Curr. Opin. Chem. Biol. 35, 1–9 (2016).

    Article  CAS  PubMed  Google Scholar 

  6. Mellmann, D., Sponholz, P., Junge, H. & Beller, M. Formic acid as a hydrogen storage material - development of homogeneous catalysts for selective hydrogen release. Chem. Soc. Rev. 45, 3954–3988 (2016).

    Article  CAS  PubMed  Google Scholar 

  7. Bulushev, D. & Ross, J. R. H. Towards sustainable production of formic acid from biomass for getting hydrogen and fuels. Chem. Sus. Chem. 11, 821–836 (2018).

    Article  CAS  Google Scholar 

  8. Reda, T., Plugge, C. M., Abram, N. J. & Hirst, J. Reversible interconversion of carbon dioxide and formate by an electroactive enzyme. Proc. Natl Acad. Sci. USA 105, 10654–10658 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Stripp, S. T. et al. Second and outer coordination sphere effects in nitrogenase, hydrogenase, formate dehydrogenase, and CO dehydrogenase. Chem. Rev. 122, 11900–11973 (2022).

  10. Schuchmann, K. & Müller, V. Autotrophy at the thermodynamic limit of life: a model for energy conservation in acetogenic bacteria. Nat. Rev. Microbiol. 12, 809–821 (2014).

    Article  CAS  PubMed  Google Scholar 

  11. Plugge, C. M., Zhang, W., Scholten, J. C. M. & Stams, A. J. M. Metabolic flexibility of sulfate-reducing bacteria. Front. Microbiol. 2, 81 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Sieber, J. R., Mcinerney, M. J. & Gunsalus, R. P. Genomic insights into syntrophy: the paradigm for anaerobic metabolic cooperation. Annu Rev. Microbiol. 66, 429–452 (2012).

    Article  CAS  PubMed  Google Scholar 

  13. Niks, D. & Hille, R. Molybdenum‐ and tungsten‐containing formate dehydrogenases and formylmethanofuran dehydrogenases: structure, mechanism, and cofactor insertion. Protein Sci. 28, 111–122 (2019).

    Article  CAS  PubMed  Google Scholar 

  14. Oliveira, A. R. et al. Toward the mechanistic understanding of enzymatic CO2 reduction. ACS Catal. 10, 3844–3856 (2020).

    Article  CAS  Google Scholar 

  15. Dietrich, H. M. et al. Membrane-anchored HDCR nanowires drive hydrogen-powered CO2 fixation. Nature 607, 823–830 (2022).

    Article  CAS  PubMed  Google Scholar 

  16. Grimaldi, S., Schoepp-Cothenet, B., Ceccaldi, P., Guigliarelli, B. & Magalon, A. The prokaryotic Mo/W-bisPGD enzymes family: a catalytic workhorse in bioenergetic. Biochim. Biophys. Acta, Bioenerg. 1827, 1048–1085 (2013).

    Article  CAS  Google Scholar 

  17. Hille, R. Molybdenum and tungsten in biology. Trends Biochem. Sci. 27, 360–367 (2002).

    Article  CAS  PubMed  Google Scholar 

  18. Schwarz, F. M., Schuchmann, K. & Müller, V. Hydrogenation of CO2 at ambient pressure catalyzed by a highly active thermostable biocatalyst. Biotechnol. Biofuels 11, 237 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  19. da Silva, S. M., Pimentel, C., Valente, F. M. A. A., Rodrigues-Pousada, C. & Pereira, I. A. C. C. Tungsten and molybdenum regulation of formate dehydrogenase expression in Desulfovibrio vulgaris Hildenborough. J. Bacteriol. 193, 2909–2916 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  20. da Silva, S. M. et al. Function of formate dehydrogenases in Desulfovibrio vulgaris Hildenborough energy metabolism. Microbiol. 159, 1760–1769 (2013).

    Article  Google Scholar 

  21. Miller, M. et al. Interfacing formate dehydrogenase with metal oxides for the reversible electrocatalysis and solar‐driven reduction of carbon dioxide. Angew. Chem. Int. Ed. 58, 4601–4605 (2019).

    Article  CAS  Google Scholar 

  22. Edwardes Moore, E. et al. A semi‐artificial photoelectrochemical tandem leaf witha CO2‐to‐formate efficiency approaching 1%. Angew. Chem. Int. Ed. 60, 26303–26307 (2021).

    Article  CAS  Google Scholar 

  23. Szczesny, J. et al. Electroenzymatic CO2 fixation using redox polymer/enzyme-modified gas diffusion electrodes. ACS Energy Lett. 5, 321–327 (2020).

    Article  CAS  Google Scholar 

  24. Alvarez-Malmagro, J. et al. Bioelectrocatalytic activity of W-formate dehydrogenase covalently immobilized on functionalized gold and graphite electrodes. ACS Appl. Mater. Interfaces 13, 11891–11900 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Antón-García, D. et al. Photoelectrochemical hybrid cell for unbiased CO2 reduction coupled to alcohol oxidation. Nat. Synth. 1, 77–86 (2022).

    Article  Google Scholar 

  26. De Bok, F. A. M. M. et al. Two W-containing formate dehydrogenases (CO2-reductases) involved in syntrophic propionate oxidation by Syntrophobacter fumaroxidans. Eur. J. Biochem. 270, 2476–2485 (2003).

    Article  PubMed  Google Scholar 

  27. Schuchmann, K. & Müller, V. Direct and reversible hydrogenation of CO2 to formate by a bacterial carbon dioxide reductase. Science 342, 1382–1385 (2013).

    Article  CAS  PubMed  Google Scholar 

  28. Raaijmakers, H. et al. Gene sequence and the 1.8 A crystal structure gene sequence and the 1.8 A of the tungsten-containing formate dehydrogenase from Desulfovibrio gigas. Structure 10, 1261–1272 (2002).

    Article  CAS  PubMed  Google Scholar 

  29. Oliveira, A. R. et al. Spectroscopic and structural characterization of reduced Desulfovibrio vulgaris Hildenborough W-FdhAB reveals stable metal coordination during catalysis. ACS Chem. Biol. 17, 1901–1909 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Vilela-Alves, G. et al. Tracking W-formate dehydrogenase structural changes during catalysis and enzyme reoxidation. Int. J. Mol. Sci. 24, 476 (2023).

    Article  CAS  Google Scholar 

  31. Grimaldi, S., Biaso, F., Burlat, B. & Guigliarelli, B. in Molybdenum and Tungsten Enzymes: Spectroscopic and Theoretical Investigations (eds Hille, R. et al.) 68–120 (The Royal Society of Chemistry, 2016).

  32. Al-Attar, S. et al. Gating of substrate access and long-range proton transfer in Escherichia coli nitrate reductase A: the essential role of a remote glutamate residue. ACS Catal. 11, 14303–14318 (2021).

    Article  CAS  Google Scholar 

  33. Iwaoka, M. & Isozumi, N. Hypervalent nonbonded interactions of a divalent sulfur atom. Implications in protein architecture and the functions. Molecules 2, 7266–7283 (2012).

    Article  Google Scholar 

  34. Léger, C. & Bertrand, P. Direct electrochemistry of redox enzymes as a tool for mechanistic studies. Chem. Rev. 108, 2379–2438 (2008).

    Article  PubMed  Google Scholar 

  35. Chiu, J. & Hogg, X. P. J. Allosteric disulfides: sophisticated molecular structures enabling flexible protein regulation. J. Biol. Chem. 294, 2949–5908 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Hartmann, T. & Leimkühler, S. The oxygen-tolerant and NAD+-dependent formate dehydrogenase from Rhodobacter capsulatus is able to catalyze the reduction of CO2 to formate. FEBS J. 280, 6083–6096 (2013).

    Article  CAS  PubMed  Google Scholar 

  37. Yu, X., Niks, D., Mulchandani, A. & Hille, R. Efficient reduction of CO2 by the molybdenum-containing formate dehydrogenase from Cupriavidus necator (Ralstonia eutropha). J. Biol. Chem. 292, 16872–16879 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Graham, J. E. et al. How a formate dehydrogenase responds to oxygen: unexpected O2 insensitivity of an enzyme harboring tungstate, selenocysteine, and [4Fe-4S] clusters. ACS Catal. 12, 10449–10471 (2022).

  39. Hogg, P. J. Disulfide bonds as switches for protein function. Trends Biochem. Sci. 28, 210–214 (2003).

    Article  CAS  PubMed  Google Scholar 

  40. Wensien, M. et al. A lysine-cysteine redox switch with an NOS bridge regulates enzyme function. Nature 593, 460–464 (2021).

    Article  CAS  PubMed  Google Scholar 

  41. Rabe von Pappenheim, F. et al. Widespread occurrence of covalent lysine–cysteine redox switches in proteins. Nat. Chem. Biol. 18, 368–375 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Schrapers, P. et al. Sulfido and cysteine ligation changes at the molybdenum cofactor during substrate conversion by formate dehydrogenase (FDH) from Rhodobacter capsulatus. Inorg. Chem. 54, 3260–3271 (2015).

    Article  CAS  PubMed  Google Scholar 

  43. Hakopian, S., Niks, D. & Hille, R. The air-inactivation of formate dehydrogenase FdsDABG from Cupriavidus necator. J. Inorg. Biochem. 231, 111788 (2022).

    Article  CAS  PubMed  Google Scholar 

  44. Agne, M., Appel, L., Seelmann, C. & Boll, M. Enoyl-coenzyme A respiration via formate cycling in syntrophic bacteria. mBio 13, 1–16 (2022).

    Article  Google Scholar 

  45. Schink, B., Montag, D., Keller, A. & Müller, N. Hydrogen or formate: alternative key players in methanogenic degradation. Environ. Microbiol Rep. 9, 189–202 (2017).

    Article  CAS  PubMed  Google Scholar 

  46. Dong, G. & Ryde, U. Reaction mechanism of formate dehydrogenase studied by computational methods. J. Biol. Inorg. Chem. 1, 1243–1254 (2018).

    Article  Google Scholar 

  47. Cypionka, H. Oxygen respiration by Desulfovibrio species. Annu. Rev. Microbiol. 54, 827–848 (2000).

    Article  CAS  PubMed  Google Scholar 

  48. Dolla, A., Fournier, M. & Dermoun, Z. Oxygen defense in sulfate-reducing bacteria. J. Biotechnol. 126, 87–100 (2006).

    Article  CAS  PubMed  Google Scholar 

  49. Mukhopadhyay, A. et al. Cell-wide responses to low-oxygen exposure in Desulfovibrio vulgaris Hildenborough. J. Bacteriol. 189, 5996–6010 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Fareleira, P. et al. Response of a strict anaerobe to oxygen: survival strategies in Desulfovibrio gigas. Microbiol. 149, 1513–1522 (2003).

    Article  CAS  Google Scholar 

  51. Imlay, J. A., Sethu, R. & Rohaun, S. K. Evolutionary adaptations that enable enzymes to tolerate oxidative stress. Free Radic. Biol. Med. 140, 4–13 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Lu, Z. & Imlay, J. A. When anaerobes encounter oxygen: mechanisms of oxygen toxicity, tolerance and defence. Nat. Rev. Microbiol. 19, 774–785 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Baumgarten, A., Redenius, I., Kranczoch, J. & Cypionka, H. Periplasmic oxygen reduction by Desulfovibrio species. Arch. Microbiol. 176, 306–309 (2001).

    Article  CAS  PubMed  Google Scholar 

  54. Ramel, F. et al. Membrane-bound oxygen reductases of the anaerobic sulfate-reducing Desulfovibrio vulgaris Hildenborough: roles in oxygen defence and electron link with periplasmic hydrogen oxidation. Microbiol. Soc. 159, 2663–2673 (2013).

  55. Fournier, M. et al. Response of the anaerobe Desulfovibrio vulgaris Hildenborough to oxidative conditions: proteome and transcript analysis. Biochimie 88, 85–94 (2006).

    Article  CAS  PubMed  Google Scholar 

  56. Vita, N., Hatchikian, E. C., Nouailler, M., Dolla, A. & Pieulle, L. Disulfide bond-dependent mechanism of protection against oxidative stress in pyruvate-ferredoxin oxidoreductase of anaerobic Desulfovibrio bacteria. Biochemistry 47, 957–964 (2008).

    Article  CAS  PubMed  Google Scholar 

  57. Rodríguez-Maciá, P. et al. Sulfide protects [FeFe] hydrogenases from O2. J. Am. Chem. Soc. 140, 9346–9350 (2018).

    Article  PubMed  Google Scholar 

  58. Marques, M. C., Coelho, R., De Lacey, A. L., Pereira, I. A. C. & Matias, P. M. The three-dimensional structure of [nifese] hydrogenase from Desulfovibrio vulgaris Hildenborough: a hydrogenase without a bridging ligand in the active site in its oxidised, ‘as-isolated’ state. J. Mol. Biol. 396, 893–907 (2010).

    Article  CAS  PubMed  Google Scholar 

  59. Felbek, C. et al. Mechanism of hydrogen sulfide-dependent inhibition of FeFe hydrogenase. ACS Catal. 11, 15162–15176 (2021).

    Article  CAS  Google Scholar 

  60. Wittenborn, E. C. et al. The solvent-exposed Fe–S D-cluster contributes to oxygen-resistance in Desulfovibrio vulgaris Ni–Fe carbon monoxide dehydrogenase. ACS Catal. 10, 7328–7335 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Keller, K. L., Wall, J. D. & Chhabra, S. in Synthetic Biology, Part A Vol. 497 (ed. Voigt, C.) 503–517 (Elsevier Inc., 2011).

  62. Meneghello, M. et al. Formate dehydrogenases reduce CO2 rather than HCO3: an electrochemical demonstration. Angew. Chem. Int. Ed. 60, 9964–9967 (2021).

    Article  CAS  Google Scholar 

  63. Fourmond, V. QSoas: a versatile software for data analysis. Anal. Chem. 88, 5050–5052 (2016).

    Article  CAS  PubMed  Google Scholar 

  64. Juanhuix, J. et al. Developments in optics and performance at BL13-XALOC, the macromolecular crystallography beamline at the ALBA Synchrotron. J. Synchrotron Radiat. 21, 679–689 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. McCarthy, A. A. et al. ID30B—a versatile beamline for macromolecular crystallography experiments at the ESRF. J. Synchrotron Radiat. 25, 1249–1260 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Nurizzo, D. et al. The ID23-1 structural biology beamline at the ESRF. J. Synchrotron Radiat. 13, 227–238 (2006).

    Article  CAS  PubMed  Google Scholar 

  67. Von Stetten, D. et al. ID30A-3 (MASSIF-3)—a beamline for macromolecular crystallography at the ESRF with a small intense beam. J. Synchrotron Radiat. 27, 844–851 (2020).

    Article  Google Scholar 

  68. Kabsch, W. XDS. Acta Crystallogr. D. Biol. Crystallogr. 66, 125–132 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Evans, P. R. & Murshudov, G. N. How good are my data and what is the resolution? Acta Crystallogr. D. Biol. Crystallogr. 69, 1204–1214 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Vonrhein, C. et al. Data processing and analysis with the autoPROC toolbox. Acta Crystallogr. D. Biol. Crystallogr. 67, 293–302 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Vonrhein, C. et al. Advances in automated data analysis and processing within autoPROC, combined with improved characterisation, mitigation and visualisation of the anisotropy of diffraction limits using STARANISO. Acta Cryst. A Found. Adv. 74, 43537 (2018).

    Article  Google Scholar 

  72. McCoy, A. J. et al. Phaser crystallographic software. J. Appl. Crystallogr. 40, 658–674 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Winn, M. D. et al. Overview of the CCP4 suite and current developments. Acta Crystallogr. D. Biol. Crystallogr. 67, 235–242 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Emsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. Features and development of Coot. Acta Crystallogr. D. Biol. Crystallogr. 66, 486–501 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Murshudov, G. N. et al. REFMAC5 for the refinement of macromolecular crystal structures. Acta Crystallogr. D. Biol. Crystallogr. 67, 355–367 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Sievers, F. & Higgins, D. G. Clustal Omega for making accurate alignments of many protein sequences. Protein Sci. 27, 135–145 (2018).

    Article  CAS  PubMed  Google Scholar 

  77. Waterhouse, A. M., Procter, J. B., Martin, D. M. A., Clamp, M. & Barton, G. J. Jalview version 2-A multiple sequence alignment editor and analysis workbench. Bioinformatics 25, 1189–1191 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Savojardo, C., Martelli, P. L., Fariselli, P. & Casadio, R. DeepSig: deep learning improves signal peptide detection in proteins. Bioinformatics 34, 1690–1696 (2018).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This work was financially supported by Fundação para a Ciência e Tecnologia (FCT, Portugal) through fellowship nos. SFRH/BD/116515/2016 (A.R.O.), DFA/BD/7897/2020 (R.R.M.) and COVID/BD/151766/2021 (A.R.O.), grant no. PTDC/BII-BBF/2050/2020 (I.A.C.P. and M.J.R.) and R&D units MOSTMICRO-ITQB (grant nos. UIDB/04612/2020 and UIDP/04612/2020) (I.A.C.P.) and UCIBIO (grant nos. UIDP/04378/2020 and UIDB/04378/2020) (M.J.R.), and Associated Laboratories LS4FUTURE (grant no. LA/P/0087/2020) (I.A.C.P.) and i4HB (grant no. LA/P/0140/2020) (M.J.R.). The European Union’s Horizon 2020 research and innovation program (grant no. 810856) is also acknowledged (I.A.C.P.). This work was also funded by the French national research agency (ANR – MOLYERE project, grant no. 16-CE-29-0010-01) (B.G.), and supported by the computing facilities of the Centre Régional de Compétences en Modélisation Moléculaire de Marseille. We thank the excellent technical assistance of João Carita from ITQB NOVA on microbial cell growth. We are also grateful to the EPR-MRS facilities of the Aix-Marseille University EPR centre and acknowledge the support of the European research infrastructure MOSBRI (grant no. 101004806) (B.G.) and the French research infrastructure INFRANALYTICS (FR2054) (B.G.). We also acknowledge the ESRF and ALBA Synchrotron for provision of synchrotron radiation facilities, and we thank the staff of the ESRF and EMBL Grenoble and ALBA for assistance and support in using beamlines ID23-1, ID30A-3, ID30B and XALOC.

Author information

Authors and Affiliations

Authors

Contributions

I.A.C.P. and A.R.O. conceived and designed biochemical experiments and sequence analysis. A.R.O. performed molecular biology experiments, protein purification, biochemical characterization, enzymatic assays, thermal shift assays, sequence analysis, EPR and electrochemical studies of WT, C845A and C872A variants and figure preparation. R.R.M. produced and characterized the M405A variant and contributed to CO2 reduction assays and figure preparation. C.M., M.J.R. and G.V.-A. designed the crystallography experiments and analyzed the crystal structures. C.M. and G.V.-A. crystallized the proteins, solved and refined all structures. G.V.-A. prepared all figures with crystal structures. K.K. helped with crystallization assays and first stages of refinement of the 8CM4 and 8CM5 structures. V.F. and C.L. designed electrochemical experiments, performed by V.F. and A.R.O. B.G. designed and analyzed EPR experiments, performed by B.G. and A.R.O. I.A.C.P., R.R.M. and N.P. designed in vivo experiments, performed by N.P. and R.R.M. I.A.C.P. and M.J.R. supervised and funded the project. A.R.O., I.A.C.P., M.J.R., C.M. and G.V.-A. wrote the manuscript with inputs from coauthors V.F., C.L. and B.G. All authors approved the final version of the manuscript.

Corresponding authors

Correspondence to Maria João Romão or Inês A. Cardoso Pereira.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Chemical Biology thanks Russ Hille and the other, anonymous reviewer(s) 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

Extended Data Fig. 1 Steady-state formate oxidation (A, C, and E) and CO2 reduction kinetics (B, D and F).

Formate oxidation (A, C, and E) and CO2 reduction (B, D and F). A and B – As-isolated WT FdhAB assays without DTT and with DTT activation, respectively. C and D – C872A variant anaerobically isolated (no DTT activation). E and F – M405A variant anaerobically isolated (no DTT activation). The concentration of enzyme used in the assays was 14 nM for A, E and F, and 0.9-2 nM for B, C and D. Data are presented as mean values ± s.d. (n = 2 or 3 assay technical replicates). Lines represent the direct fitting of Michaelis-Menten equation to experimental data.

Source data

Extended Data Fig. 2 EPR spectra of WV species in WT FdhAB enzyme and in C872A and M405A variants.

Experimental spectra are in black, and simulations are shown below in red. a) resting WT sample poised at −468 mV by reduction with dithionite; b) DTT activated WT sample poised at −395 mV by reduction with formate; c) C872A sample poised at −443 mV by reduction with formate; d) C872A sample poised at −469 mV by reduction with dithionite; e) M405A reduced by dithionite. EPR conditions: temperature 80 K; microwave power 40 mW at 9.479 GHz; modulation amplitude 1 mT at 100 kHz. Simulated spectra result from the superimposition of WVF and WVD signals with parameters given in Table S2 with the following WVD /WVF ratio: a) 90/10; b) 5/95; c) 10/90; d) 80/20. For the sake of clarity, the radical signal at g=2.00 arising from mediators in redox titrations was not simulated.

Extended Data Fig. 3 FdhAB W active site.

FdhAB W active site in aerobic C872A_ox crystal structure (a, b, c) and C872A_anox anaerobic one (d). The results for three different labile group refinement strategies of the labile ligand, for the C872A-O2 model, are presented: (a) refinement with an oxygen replacing the sulfido ligand; (b) refinement with a sulfido ligand with an occupancy of 1; and (c) refinement with a sulfido ligand refined with an occupancy of 0.556. In (d) the W center for the C872Aanoxic-CHNH2O is shown. The pterin domains of both MGD cofactors and the sidechain of U192 are represented as light blue or blue sticks, respectively, for C872A-O2 and C872Aanoxic-CHNH2O, the W ion as a blue sphere and the sulfido ligand by a yellow stick. The 2Fo-Fc electron density map for each refinement is shown as a magenta mesh, at 1σ, and the Fo-Fc map is represented as a green mesh, for positive density, and as a red mesh, for negative density, both at 3σ.

Extended Data Fig. 4 Comparison of the active site in WT and C872A_anox structures.

Superposition of FdhAB WT (gray) and C872A_anox structures (blue) showing the new H193 conformation and respective hydrogen bonding network. The W ion is represented as a violet sphere and its two bound MGDs are shown as sticks. Distances are in Å.

Extended Data Fig. 5 A formamide molecule is present near the active site (C872A_anox structure) in the formate channel.

R441 and T450 are hydrogen bonding the formamide ligand. The W ion is represented as a violet sphere and its two bound MGDs are shown as sticks. The 2Fo-Fc electron density map is shown as a magenta mesh, at 1σ. Distances are in Å.

Extended Data Fig. 6 EPR spectrum of WV in reduced M405A variant.

a) M405A sample as-isolated anaerobically; b) after reduction with dithionite; c) after reduction with formate; d) formate-reduced sample after washing-off formate. EPR conditions: temperature 80 K; microwave power 40 mW; modulation amplitude 1 mT at 100 kHz.

Source data

Extended Data Fig. 7 Active site environment of the M405A variant.

Superposition of WT (grey) and M405A variant (yellow) is shown. Black arrows highlight the conformational changes. The W ion is represented as a violet sphere and its two bound MGDs are shown as sticks.

Extended Data Fig. 8 Rate of O2 inactivation.

Solid lines show the derivative of the logarithm of the current for a– DTT-activated WT FdhAB, b– Anaerobically purified WT FdhAB, c- C845A and d- C872A variants, after injection of 30 µM of O2. Dashed lines are the representative fits of the kinetic model in equation [3] (Supplementary Information).

Source data

Supplementary information

Supplementary Information

Supplementary Data 1–3, Figs. 1–3, Tables 1–7 and references.

Reporting Summary

Supplementary Data 1

Source Data for Supplementary Fig. 2.

Supplementary Data 2

Source Data for Supplementary Fig. 3.

Source data

Source Data Fig. 1

Statistical source data for Fig. 1.

Source Data Fig. 2

Source data for Fig. 2.

Source Data Fig. 4

Statistical source data for Fig. 4.

Source Data Fig. 5

Statistical source data for Fig. 5.

Source Data Table 1

Statistical source data for Table 1.

Source Data Extended Data Fig. 1

Statistical source data for Extended Data Fig. 1.

Source Data Extended Data Fig. 6

Source data for Extended Data Fig. 6.

Source Data Extended Data Fig. 8

Source data for Extended Data Fig. 8.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Oliveira, A.R., Mota, C., Vilela-Alves, G. et al. An allosteric redox switch involved in oxygen protection in a CO2 reductase. Nat Chem Biol 20, 111–119 (2024). https://doi.org/10.1038/s41589-023-01484-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41589-023-01484-2

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