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:

Trace gas oxidizers are widespread and active members of soil microbial communities

Abstract

Soil microorganisms globally are thought to be sustained primarily by organic carbon sources. Certain bacteria also consume inorganic energy sources such as trace gases, but they are presumed to be rare community members, except within some oligotrophic soils. Here we combined metagenomic, biogeochemical and modelling approaches to determine how soil microbial communities meet energy and carbon needs. Analysis of 40 metagenomes and 757 derived genomes indicated that over 70% of soil bacterial taxa encode enzymes to consume inorganic energy sources. Bacteria from 19 phyla encoded enzymes to use the trace gases hydrogen and carbon monoxide as supplemental electron donors for aerobic respiration. In addition, we identified a fourth phylum (Gemmatimonadota) potentially capable of aerobic methanotrophy. Consistent with the metagenomic profiling, communities within soil profiles from diverse habitats rapidly oxidized hydrogen, carbon monoxide and to a lesser extent methane below atmospheric concentrations. Thermodynamic modelling indicated that the power generated by oxidation of these three gases is sufficient to meet the maintenance needs of the bacterial cells capable of consuming them. Diverse bacteria also encode enzymes to use trace gases as electron donors to support carbon fixation. Altogether, these findings indicate that trace gas oxidation confers a major selective advantage in soil ecosystems, where availability of preferred organic substrates limits microbial growth. The observation that inorganic energy sources may sustain most soil bacteria also has broad implications for understanding atmospheric chemistry and microbial biodiversity in a changing world.

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: Energy conservation and carbon acquisition strategies of global soil bacteria.
Fig. 2: Radial maximum-likelihood phylogenetic trees showing the sequence diversity and taxonomic distribution of key enzymes associated with trace gas oxidation.
Fig. 3: Measurement of oxidation of the trace gases H2, CO and CH4 across four Australian soil ecosystems.

Similar content being viewed by others

Data availability

All data supporting the findings of the present study are available. All numerical data used to make figures are provided in Source Data 1 (XLSX format) and all raw phylogenetic trees are provided in Newick (NWK) format in Source Data 2 (ZIP format). All metagenomes sequenced for this project can be accessed at the Sequence Read Archive with accession number PRJNA656125 (https://www.ncbi.nlm.nih.gov/bioproject/PRJNA656125) and all MAGs are available at FigShare (https://figshare.com/articles/dataset/Metagenome_Assembled_Genomes/12782543). All previously sequenced metagenomes and metatranscriptomes analysed in this study are available at NCBI BioProject with the accession numbers listed in Supplementary Table 2. Source data are provided with this paper.

References

  1. Fierer, N. Embracing the unknown: disentangling the complexities of the soil microbiome. Nat. Rev. Microbiol. 15, 579–590 (2017).

    CAS  PubMed  Google Scholar 

  2. Bell, T., Newman, J. A., Silverman, B. W., Turner, S. L. & Lilley, A. K. The contribution of species richness and composition to bacterial services. Nature 436, 1157 (2005).

    CAS  PubMed  Google Scholar 

  3. Delgado-Baquerizo, M. et al. A global atlas of the dominant bacteria found in soil. Science 359, 320–325 (2018).

    CAS  PubMed  Google Scholar 

  4. Bahram, M. et al. Structure and function of the global topsoil microbiome. Nature 560, 233–237 (2018).

    CAS  PubMed  Google Scholar 

  5. Jones, S. E. & Lennon, J. T. Dormancy contributes to the maintenance of microbial diversity. Proc. Natl Acad. Sci. USA 107, 5881–5886 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Lennon, J. T. & Jones, S. E. Microbial seed banks: the ecological and evolutionary implications of dormancy. Nat. Rev. Microbiol. 9, 119–130 (2011).

    CAS  PubMed  Google Scholar 

  7. Schimel, J. & Schaeffer, S. M. Microbial control over carbon cycling in soil. Front. Microbiol. 3, 348 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Janssen, P. H. Identifying the dominant soil bacterial taxa in libraries of 16S rRNA and 16S rRNA genes. Appl. Environ. Microbiol. 72, 1719–1728 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Joseph, S. J., Hugenholtz, P., Sangwan, P., Osborne, C. A. & Janssen, P. H. Laboratory cultivation of widespread and previously uncultured soil bacteria. Appl. Environ. Microbiol. 69, 7210–7215 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Conrad, R. Soil microorganisms as controllers of atmospheric trace gases (H2, CO, CH4, OCS, N2O, and NO). Microbiol. Rev. 60, 609–640 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Beeckman, F., Motte, H. & Beeckman, T. Nitrification in agricultural soils: impact, actors and mitigation. Curr. Opin. Biotechnol. 50, 166–173 (2018).

    CAS  PubMed  Google Scholar 

  12. Khdhiri, M. et al. Soil carbon content and relative abundance of high affinity H2-oxidizing bacteria predict atmospheric H2 soil uptake activity better than soil microbial community composition. Soil Biol. Biochem. 85, 1–9 (2015).

    CAS  Google Scholar 

  13. Ehhalt, D. H. & Rohrer, F. The tropospheric cycle of H2: a critical review. Tellus B. 61, 500–535 (2009).

    Google Scholar 

  14. Novelli, P. C., Masarie, K. A., Tans, P. P. & Lang, P. M. Recent changes in atmospheric carbon monoxide. Science 263, 1587–1590 (1994).

    CAS  PubMed  Google Scholar 

  15. Novelli, P. C., Masarie, K. A. & Lang, P. M. Distributions and recent changes of carbon monoxide in the lower troposphere. J. Geophys. Res. Atmos. 103, 19015–19033 (1998).

    CAS  Google Scholar 

  16. Kirschke, S. et al. Three decades of global methane sources and sinks. Nat. Geosci. 6, 813–823 (2013).

    CAS  Google Scholar 

  17. Constant, P., Poissant, L. & Villemur, R. Tropospheric H2 budget and the response of its soil uptake under the changing environment. Sci. Total Environ. 407, 1809–1823 (2009).

    CAS  PubMed  Google Scholar 

  18. Berney, M., Greening, C., Conrad, R., Jacobs, W. R. & Cook, G. M. An obligately aerobic soil bacterium activates fermentative hydrogen production to survive reductive stress during hypoxia. Proc. Natl Acad. Sci. USA 111, 11479–11484 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Tarr, M. A., Miller, W. L. & Zepp, R. G. Direct carbon monoxide photoproduction from plant matter. J. Geophys. Res. Atmos. 100, 11403–11413 (1995).

    CAS  Google Scholar 

  20. King, G. M. & Weber, C. F. Distribution, diversity and ecology of aerobic CO-oxidizing bacteria. Nat. Rev. Microbiol. 5, 107–118 (2007).

    CAS  PubMed  Google Scholar 

  21. Conrad, R. The global methane cycle: recent advances in understanding the microbial processes involved. Environ. Microbiol. Rep. 1, 285–292 (2009).

    CAS  PubMed  Google Scholar 

  22. Constant, P., Chowdhury, S. P., Pratscher, J. & Conrad, R. Streptomycetes contributing to atmospheric molecular hydrogen soil uptake are widespread and encode a putative high-affinity [NiFe]-hydrogenase. Environ. Microbiol. 12, 821–829 (2010).

    CAS  PubMed  Google Scholar 

  23. Greening, C., Berney, M., Hards, K., Cook, G. M. & Conrad, R. A soil actinobacterium scavenges atmospheric H2 using two membrane-associated, oxygen-dependent [NiFe] hydrogenases. Proc. Natl Acad. Sci. USA 111, 4257–4261 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Cordero, P. R. F. et al. Atmospheric carbon monoxide oxidation is a widespread mechanism supporting microbial survival. ISME J. 13, 2868–2881 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Mörsdorf, G., Frunzke, K., Gadkari, D. & Meyer, O. Microbial growth on carbon monoxide. Biodegradation 3, 61–82 (1992).

    Google Scholar 

  26. Greening, C. et al. Genomic and metagenomic surveys of hydrogenase distribution indicate H2 is a widely utilised energy source for microbial growth and survival. ISME J. 10, 761–777 (2016).

    CAS  PubMed  Google Scholar 

  27. Tveit, A. T. et al. Widespread soil bacterium that oxidizes atmospheric methane. Proc. Natl Acad. Sci. USA 116, 8515–8524 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Islam, Z. F. et al. A widely distributed hydrogenase oxidises atmospheric H2 during bacterial growth. ISME J. https://doi.org/10.1038/s41396-020-0713-4 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  29. King, G. M. Molecular and culture-based analyses of aerobic carbon monoxide oxidizer diversity. Appl. Environ. Microbiol. 69, 7257–7265 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Greening, C. et al. Persistence of the dominant soil phylum Acidobacteria by trace gas scavenging. Proc. Natl Acad. Sci. USA 112, 10497–10502 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Berney, M. & Cook, G. M. Unique flexibility in energy metabolism allows mycobacteria to combat starvation and hypoxia. PLoS ONE 5, e8614 (2010).

    PubMed  PubMed Central  Google Scholar 

  32. Greening, C., Villas-Bôas, S. G., Robson, J. R., Berney, M. & Cook, G. M. The growth and survival of Mycobacterium smegmatis is enhanced by co-metabolism of atmospheric H2. PLoS ONE 9, e103034 (2014).

    PubMed  PubMed Central  Google Scholar 

  33. Liot, Q. & Constant, P. Breathing air to save energy – new insights into the ecophysiological role of high-affinity [NiFe]-hydrogenase in Streptomyces avermitilis. Microbiologyopen 5, 47–59 (2016).

    CAS  PubMed  Google Scholar 

  34. Greening, C., Grinter, R. & Chiri, E. Uncovering the metabolic strategies of the dormant microbial majority: towards integrative approaches. mSystems 4, e00107-19 (2019).

    PubMed  PubMed Central  Google Scholar 

  35. Khdhiri, M., Piché-Choquette, S., Tremblay, J., Tringe, S. G. & Constant, P. Meta-omics survey of [NiFe]-hydrogenase genes fails to capture drastic variations in H2-oxidation activity measured in three soils exposed to H2. Soil Biol. Biochem. 125, 239–243 (2018).

    CAS  Google Scholar 

  36. Quiza, L., Lalonde, I., Guertin, C. & Constant, P. Land-use influences the distribution and activity of high affinity CO-oxidizing bacteria associated to type I-coxL genotype in soil. Front. Microbiol. 5, 271 (2014).

    PubMed  PubMed Central  Google Scholar 

  37. Ji, M. et al. Atmospheric trace gases support primary production in Antarctic desert surface soil. Nature 552, 400–403 (2017).

    CAS  PubMed  Google Scholar 

  38. Lynch, R. C., Darcy, J. L., Kane, N. C., Nemergut, D. R. & Schmidt, S. K. Metagenomic evidence for metabolism of trace atmospheric gases by high-elevation desert Actinobacteria. Front. Microbiol. 5, 698 (2014).

    PubMed  PubMed Central  Google Scholar 

  39. King, G. M. Contributions of atmospheric CO and hydrogen uptake to microbial dynamics on recent Hawaiian volcanic deposits. Appl. Environ. Microbiol. 69, 4067–4075 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Constant, P., Chowdhury, S. P., Hesse, L., Pratscher, J. & Conrad, R. Genome data mining and soil survey for the novel Group 5 [NiFe]-hydrogenase to explore the diversity and ecological importance of presumptive high-affinity H2-oxidizing bacteria. Appl. Environ. Microbiol. 77, 6027–6035 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Pandelia, M. E., Lubitz, W. & Nitschke, W. Evolution and diversification of Group 1 [NiFe] hydrogenases. Is there a phylogenetic marker for O2-tolerance? Biochim. Biophys. Acta - Bioenerg. 1817, 1565–1575 (2012).

    CAS  Google Scholar 

  42. Myers, M. R. & King, G. M. Isolation and characterization of Acidobacterium ailaaui sp. nov., a novel member of Acidobacteria subdivision 1, from a geothermally heated Hawaiian microbial mat. Int. J. Syst. Evol. Microbiol. 66, 5328–5335 (2016).

    CAS  PubMed  Google Scholar 

  43. Cunliffe, M. Correlating carbon monoxide oxidation with cox genes in the abundant marine Roseobacter clade. ISME J. 5, 685–691 (2011).

    CAS  PubMed  Google Scholar 

  44. van Kessel, M. A. H. J. et al. Complete nitrification by a single microorganism. Nature 528, 555–559 (2015).

    PubMed  PubMed Central  Google Scholar 

  45. Daims, H. et al. Complete nitrification by Nitrospira bacteria. Nature 528, 504–509 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Knief, C. Diversity and habitat preferences of cultivated and uncultivated aerobic methanotrophic bacteria evaluated based on pmoA as molecular marker. Front. Microbiol. 6, 1346 (2015).

    PubMed  PubMed Central  Google Scholar 

  47. Chiri, E. et al. Termite mounds contain soil-derived methanotroph communities kinetically adapted to elevated methane concentrations. ISME J. 14, 2715–2731 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Dunfield, P. F. et al. Methane oxidation by an extremely acidophilic bacterium of the phylum Verrucomicrobia. Nature 450, 879–882 (2007).

    CAS  PubMed  Google Scholar 

  49. Haroon, M. F. et al. Anaerobic oxidation of methane coupled to nitrate reduction in a novel archaeal lineage. Nature 500, 567–570 (2013).

    CAS  PubMed  Google Scholar 

  50. Khadka, R. et al. Evolutionary history of copper membrane monooxygenases. Front. Microbiol. 9, 2493 (2018).

    PubMed  PubMed Central  Google Scholar 

  51. Van Asperen, H. et al. The role of photo-and thermal degradation for CO2 and CO fluxes in an arid ecosystem. Biogeosciences 12, 4161–4174 (2015).

    Google Scholar 

  52. Moxley, J. M. & Smith, K. A. Carbon monoxide production and emission by some Scottish soils. Tellus B. Chem. Phys. Meteorol. 50, 151–162 (1998).

    Google Scholar 

  53. Dutaur, L. & Verchot, L. V. A global inventory of the soil CH4 sink. Global Biogeochem. Cycles 21, GB4013 (2007).

    Google Scholar 

  54. McLearn, N. & Dong, Z. Microbial nature of the hydrogen-oxidizing agent in hydrogen-treated soil. Biol. Fertil. Soils 35, 465–469 (2002).

    CAS  Google Scholar 

  55. King, G. M. Attributes of atmospheric carbon monoxide oxidation by Maine forest soils. Appl. Environ. Microbiol. 65, 5257–5264 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Kolb, S., Knief, C., Dunfield, P. F. & Conrad, R. Abundance and activity of uncultured methanotrophic bacteria involved in the consumption of atmospheric methane in two forest soils. Environ. Microbiol. 7, 1150–1161 (2005).

    CAS  PubMed  Google Scholar 

  57. Priemé, A. & Christensen, S. Seasonal and spatial variation of methane oxidation in a Danish spruce forest. Soil Biol. Biochem. 29, 1165–1172 (1997).

    Google Scholar 

  58. Chen, Q., Popa, M. E., Batenburg, A. M. & Röckmann, T. Isotopic signatures of production and uptake of H2 by soil. Atmos. Chem. Phys. 15, 13003–13021 (2015).

    CAS  Google Scholar 

  59. DeLong, J. P., Okie, J. G., Moses, M. E., Sibly, R. M. & Brown, J. H. Shifts in metabolic scaling, production, and efficiency across major evolutionary transitions of life. Proc. Natl Acad. Sci. USA 107, 12941–12945 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Kempes, C. P. et al. Drivers of bacterial maintenance and minimal energy requirements. Front. Microbiol. 8, 31 (2017).

    PubMed  PubMed Central  Google Scholar 

  61. Marschall, E., Jogler, M., Henßge, U. & Overmann, J. Large‐scale distribution and activity patterns of an extremely low‐light‐adapted population of green sulfur bacteria in the Black Sea. Environ. Microbiol. 12, 1348–1362 (2010).

    CAS  PubMed  Google Scholar 

  62. Tijhuis, L., Van Loosdrecht, M. C. & Heijnen, J. J. A thermodynamically based correlation for maintenance gibbs energy requirements in aerobic and anaerobic chemotrophic growth. Biotechnol. Bioeng. 42, 509–519 (1993).

    CAS  PubMed  Google Scholar 

  63. LaRowe, D. E. & Amend, J. P. Power limits for microbial life. Front. Microbiol. 6, 718 (2015).

    PubMed  PubMed Central  Google Scholar 

  64. Bradley, J. A. et al. Widespread energy limitation to life in global subseafloor sediments. Sci. Adv. 6, eaba0697 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Røy, H. et al. Aerobic microbial respiration in 86-million-year-old deep-sea red clay. Science 336, 922–925 (2012).

    PubMed  Google Scholar 

  66. Conrad, R. in Current Perspectives in Microbial Ecology (eds Klug, M. G. & Reddy, C. A.) 461–467 (American Society for Microbiology, 1984).

  67. Conrad, R. Soil microorganisms oxidizing atmospheric trace gases (CH4, CO, H2, NO). Indian J. Microbiol. 39, 193–203 (1999).

    Google Scholar 

  68. Meredith, L. K. et al. Ecosystem fluxes of hydrogen in a mid‐latitude forest driven by soil microorganisms and plants. Glob. Change Biol. 23, 906–919 (2017).

    Google Scholar 

  69. Carini, P. et al. Relic DNA is abundant in soil and obscures estimates of soil microbial diversity. Nat. Microbiol. 2, 16242 (2016).

    PubMed  Google Scholar 

  70. Schmitz, R. A. et al. The thermoacidophilic methanotroph Methylacidiphilum fumariolicum SolV oxidizes subatmospheric H2 with a high-affinity, membrane-associated [NiFe] hydrogenase. ISME J. 14, 1223–1232 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Islam, Z. F. et al. Two Chloroflexi classes independently evolved the ability to persist on atmospheric hydrogen and carbon monoxide. ISME J. 13, 1801–1813 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  72. Hüppi, R. et al. Restricting the nonlinearity parameter in soil greenhouse gas flux calculation for more reliable flux estimates. PLoS ONE https://doi.org/10.1371/journal.pone.0200876 (2018).

  73. Caporaso, J. G. et al. Global patterns of 16S rRNA diversity at a depth of millions of sequences per sample. Proc. Natl Acad. Sci. USA 108, 4516–4522 (2011).

    CAS  PubMed  Google Scholar 

  74. Uritskiy, G. V., DiRuggiero, J. & Taylor, J. MetaWRAP—a flexible pipeline for genome-resolved metagenomic data analysis. Microbiome 6, 158 (2018).

    PubMed  PubMed Central  Google Scholar 

  75. Li, D., Liu, C.-M., Luo, R., Sadakane, K. & Lam, T.-W. MEGAHIT: an ultra-fast single-node solution for large and complex metagenomics assembly via succinct de Bruijn graph. Bioinformatics 31, 1674–1676 (2015).

    CAS  PubMed  Google Scholar 

  76. Nurk, S., Meleshko, D., Korobeynikov, A. & Pevzner, P. A. metaSPAdes: a new versatile metagenomic assembler. Genome Res. 27, 824–834 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  77. Kang, D. D., Froula, J., Egan, R. & Wang, Z. MetaBAT, an efficient tool for accurately reconstructing single genomes from complex microbial communities. Peer J 3, e1165 (2015).

    PubMed  PubMed Central  Google Scholar 

  78. Kang, D. et al. MetaBAT 2: an adaptive binning algorithm for robust and efficient genome reconstruction from metagenome assemblies. Peer J 7, e7359 (2019).

    PubMed  PubMed Central  Google Scholar 

  79. Wu, Y.-W., Simmons, B. A. & Singer, S. W. MaxBin 2.0: an automated binning algorithm to recover genomes from multiple metagenomic datasets. Bioinformatics 32, 605–607 (2015).

    PubMed  Google Scholar 

  80. Olm, M. R., Brown, C. T., Brooks, B. & Banfield, J. F. dRep: a tool for fast and accurate genomic comparisons that enables improved genome recovery from metagenomes through de-replication. ISME J. 11, 2864–2868 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  81. Parks, D. H., Imelfort, M., Skennerton, C. T., Hugenholtz, P. & Tyson, G. W. CheckM: assessing the quality of microbial genomes recovered from isolates, single cells, and metagenomes. Genome Res. 25, 1043–1055 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  82. Bowers, R. M. et al. Minimum information about a single amplified genome (MISAG) and a metagenome-assembled genome (MIMAG) of bacteria and archaea. Nat. Biotechnol. 35, 725–731 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  83. Parks, D. H. et al. A standardized bacterial taxonomy based on genome phylogeny substantially revises the tree of life. Nat. Biotechnol. 36, 996–1004 (2018).

    CAS  PubMed  Google Scholar 

  84. Chaumeil, P.-A., Mussig, A. J., Hugenholtz, P. & Parks, D. H. GTDB-Tk: a toolkit to classify genomes with the Genome Taxonomy Database. Bioinformatics 36, 1925–1927 (2020).

    CAS  Google Scholar 

  85. Markowitz, V. M. et al. IMG: the integrated microbial genomes database and comparative analysis system. Nucleic Acids Res. 40, D115–D122 (2012).

    CAS  PubMed  Google Scholar 

  86. Leinonen, R., Sugawara, H. & Shumway, M., Collaboration, I. N. S. D. The sequence read archive. Nucleic Acids Res. 39, D19–D21 (2010).

    PubMed  PubMed Central  Google Scholar 

  87. Jansson, J. K. Next generation ecosystem experiment (NGEE) in the Arctic https://doi.org/10.25585/1487743 (2013).

  88. Constant, P. Metagenomic and Metatranscriptomic Analysis of Soil Biogeochemical Processes Sustained by Interspecific Transfer of Molecular Hydrogen (JGI, 2014); https://doi.org/10.25585/1487738

  89. Tiedje, J. M. Metagenomic Analysis of the Rhizosphere of Three Biofuel Crops at the KBS Intensive Site (JGI, 2013); https://doi.org/10.25585/1488010

  90. Mohn, W. W. Metagenomic and Metatranscriptomic Analysis of Forest Soil Communities Across North America (JGI, 2012); https://doi.org/10.25585/1487719

  91. Kalyuzhnaya, M. G. Systems Level Insights into Methane Cycling in Arid and Semi-Arid Ecosystems via Community Metagenomics and Metatranscriptomics (JGI, 2016); https://doi.org/10.25585/1488146

  92. Cadillo-Quiroz, H. Microbial Composition and Metagenomic Functional Potential Across Tropical Peatlands: Comparative Evaluation and Modeling of C decomposition to Greenhouse Gases (JGI, 2015); https://doi.org/10.25585/1488117

  93. Oliveira, R. Unravelling Microbial Communities Associated with Native Plant Species from P-Impoverished Soils of a Global Biodiversity Hotspot (JGI, 2017); https://doi.org/10.25585/1488283

  94. Pett-Ridge, J. Microbial Carbon Transformations in Wet Tropical Soils: Effects of Redox Fluctuation (JGI, 2017); https://doi.org/10.25585/1488160

  95. Hyatt, D. et al. Prodigal: prokaryotic gene recognition and translation initiation site identification. BMC Bioinform. 11, 119 (2010).

    Google Scholar 

  96. Buchfink, B., Xie, C. & Huson, D. H. Fast and sensitive protein alignment using DIAMOND. Nat. Methods 12, 59–60 (2014).

    PubMed  Google Scholar 

  97. Anantharaman, K. et al. Thousands of microbial genomes shed light on interconnected biogeochemical processes in an aquifer system. Nat. Commun. 7, 13219 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  98. Altschul, S. F., Gish, W., Miller, W., Myers, E. W. & Lipman, D. J. Basic local alignment search tool. J. Mol. Biol. 215, 403–410 (1990).

    CAS  PubMed  Google Scholar 

  99. Kopylova, E., Noé, L. & Touzet, H. SortMeRNA: fast and accurate filtering of ribosomal RNAs in metatranscriptomic data. Bioinform. 28, 3211–3217 (2012).

    CAS  Google Scholar 

  100. Patro, R., Duggal, G., Love, M. I., Irizarry, R. A. & Kingsford, C. Salmon provides fast and bias-aware quantification of transcript expression. Nat. Methods 14, 417–419 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  101. Fu, L., Niu, B., Zhu, Z., Wu, S. & Li, W. CD-HIT: accelerated for clustering the next-generation sequencing data. Bioinform. 28, 3150–3152 (2012).

    CAS  Google Scholar 

  102. Kumar, S., Stecher, G. & Tamura, K. MEGA7: Molecular Evolutionary Genetics Analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 33, 1870–1874 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  103. Kelley, L. A., Mezulis, S., Yates, C. M., Wass, M. N. & Sternberg, M. J. E. The Phyre2 web portal for protein modeling, prediction and analysis. Nat. Protoc. 10, 845–858 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  104. Smith, S. M. et al. Crystal structure and characterization of particulate methane monooxygenase from Methylocystis species strain M. Biochemistry 50, 10231–10240 (2011).

    CAS  PubMed  Google Scholar 

  105. McMurdie, P. J. & Holmes, S. phyloseq: an R package for reproducible interactive analysis and graphics of microbiome census data. PLoS ONE 8, e61217 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  106. Dixon, P. VEGAN, a package of R functions for community ecology. J. Veg. Sci. 14, 927–930 (2003).

    Google Scholar 

  107. Helgeson, H., Kirkham, D. & Flowers, G. Theoretical prediction of the thermodynamic behavior of aqueous electrolytes by high pressures and temperatures; IV, calculation of activity coefficients, osmotic coefficients, and apparent molal and standard and relative partial molal properties to 600 d. Am. J. Sci. 281, 1249–1516 (1981).

    CAS  Google Scholar 

  108. Shock, E. L. & Helgeson, H. C. Calculation of the thermodynamic and transport properties of aqueous species at high pressures and temperatures: correlation algorithms for ionic species and equation of state predictions to 5 kb and 1000 °C. Geochim. Cosmochim. Acta 52, 2009–2036 (1988).

    CAS  Google Scholar 

  109. Shock, E. L., Oelkers, E. H., Johnson, J. W., Sverjensky, D. A. & Helgeson, H. C. Calculation of the thermodynamic properties of aqueous species at high pressures and temperatures. Effective electrostatic radii, dissociation constants and standard partial molal properties to 1000 °C and 5 kbar. J. Chem. Soc. Faraday Trans. 88, 803–826 (1992).

    CAS  Google Scholar 

  110. Johnson, J. W., Oelkers, E. H. & Helgeson, H. C. A software package for calculating the standard molal thermodynamic properties of minerals, gases, aqueous species and reactions from 1 to 5000 bars and 0° to 1000 °C. Comput. Geosci. 18, 899–947 (1992).

    Google Scholar 

  111. Shock, E. L., Helgeson, H. C. & Sverjensky, D. A. Calculation of the thermodynamic and transport properties of aqueous species at high pressures and temperatures: standard partial molal properties of inorganic neutral species. Geochim. Cosmochim. Acta 53, 2157–2183 (1989).

    CAS  Google Scholar 

  112. Shock, E. L. & Helgeson, H. C. Calculation of the thermodynamic and transport properties of aqueous species at high pressures and temperatures: standard partial molal properties of organic species. Geochim. Cosmochim. Acta 54, 915–945 (1990).

    CAS  Google Scholar 

  113. Sverjensky, D. A., Shock, E. L. & Helgeson, H. C. Prediction of the thermodynamic properties of aqueous metal complexes to 1000 °C and 5 kb. Geochim. Cosmochim. Acta 61, 1359–1412 (1997).

    CAS  PubMed  Google Scholar 

  114. Schulte, M. D., Shock, E. L. & Wood, R. H. The temperature dependence of the standard-state thermodynamic properties of aqueous nonelectrolytes. Geochim. Cosmochim. Acta 65, 3919–3930 (2001).

    CAS  Google Scholar 

  115. Amend, J. P. & LaRowe, D. E. Minireview: demystifying microbial reaction energetics. Environ. Microbiol. 21, 3539–3547 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  116. Větrovský, T. & Baldrian, P. The variability of the 16S rRNA gene in bacterial genomes and its consequences for bacterial community analyses. PLoS ONE 8, e57923 (2013).

    PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This research was supported by an Australian Research Council DECRA Fellowship (DE170100310; C.G.), a Discovery Project grant (DP180101762; C.G. and P.L.M.C.), a Swiss National Foundation Early Mobility Postdoctoral Fellowship (P2EZP3_178421; E.C.), a NSFC grant (41906076; X.D.), a NERC grant (NE/T010967/1; J.B.), a Humboldt Foundation Fellowship (J.B.), Monash University PhD scholarships (S.K.B. and P.M.L.), a Holsworth Wildlife Research Endowment Grant (S.K.B) and an NHMRC EL2 Fellowship (APP1178715; salary for C.G.). We thank M. Chuvochina for etymological advice, J. Lloyd for field assistance and W. Whitman, S. Chown, S. Morales, M. Stott and G. Cook for formative discussions.

Author information

Authors and Affiliations

Authors

Contributions

C.G. conceived and supervised this study. S.K.B., C.G., E.C. and P.A.N. designed experiments and analysed data. Different authors were responsible for performing fieldwork (S.K.B., E.C., P.M.L., C.G.), laboratory work (S.K.B., T.J., E.C.), metagenome assembly (X.D.), metagenome analysis (S.K.B., C.G., X.D., E.C.), phylogenetic analysis (C.G.), protein modelling (R.G.) and thermodynamic modelling (J.A.B., P.A.N., D.L., E.C., S.K.B.). S.K.A. and P.L.M.C. provided logistical and theoretical support. C.G., S.K.B. and P.A.N. wrote the paper with input from all authors.

Corresponding authors

Correspondence to Eleonora Chiri or Chris Greening.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Peer reviewer reports are available.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 Composition of the bacterial and archaeal communities sequenced in each soil metagenome.

Stacked barcharts depicting the relative abundance of different phyla in (a) Australian soils and (b) global soils based on reads for the single-copy ribosomal protein gene rplP. Alpha diversity is shown as observed and estimated richness (Chao1) in (c) Australian soils and (d) global soils. Boxplots show median, lower and upper quartile, and minimum and maximum values. This is based on four biological replicates for the four Australian soils (n = 16) at depths of 0-5, 5-10, 15-20, and 25-30 cm, and biological triplicates for the eight global soils (n = 24). For beta diversity, abundance-based distance matrix Bray-Curtis diversity is visualized on a multidimensional PcoA plot for both Australian and global soils. Testing for significant differences in community structure between depths and ecosystems was performed using a one-way PERMANOVA with 999 permutations.

Source data

Extended Data Fig. 2 Metatranscriptomic analysis comparing metabolic marker gene expression across three global soils with paired metagenome and metatranscriptome datasets.

(a) Relative abundance of quality-filtered metatranscriptomic short reads expressed as the total reads per kilobase per million (RPKM). (b) Relative abundance of genes in each metatranscriptome mapped to the corresponding assemblies from each soil expressed as total transcripts per million (TPM).

Source data

Extended Data Fig. 3 Analysis of aerobic methane oxidation enzymes and pathways in sampled soils.

(a) Maximum-likelihood tree of amino acid sequences of particulate methane monooxygenase A subunit (PmoA), a marker for aerobic methane oxidation. The tree shows 2 sequences from soil metagenome-assembled genomes (blue) and 12 sequences from unbinned contigs (red) alongside 93 representative sequences from NCBI reference genomes (black). The tree shows the affiliation of the PmoA from Candidatus Methylotropicum kingii with those of amplicons of the tropical upland soil cluster (TUSC). Also shown are reference sequences of other members of the copper membrane monooxygenase superfamily, namely ammonia monooxygenase (AmoA), hydrocarbon monooxygenase (HmoA), and groups of unknown function (including PxmA). The tree was constructed using the JTT matrix-based model, used all sites, and was bootstrapped with 50 replicates and midpoint-rooted. Source data is provided in Newick (nwk) format. (b) Metabolic reconstruction of the putative novel methanotroph Candidatus Methylotropicum kingii. The core pathways associated with energy conservation and carbon acquisition are shown, with genes detected shown in italics. The bacterium is predicted to use methane, methanol, and acetate as energy and carbon sources. In addition, it can use molecular hydrogen as an electron donor via a group 1f [NiFe]-hydrogenase. The bacterium is predicted to use the electron acceptors oxygen via a cytochrome c oxidase and nitrous oxide via a nitrous oxide reductase. Its particulate methane monooxygenase forms a distinct phylogenetic lineage with amplicons from the Tropical Upland Soil Cluster (TUSC), whereas its methanol dehydrogenase is closely related to those in previously sequenced Gemmatimonadota MAGs inferred to be methylotrophic. The genome encodes key enzymes for the serine cycle for assimilation of one-carbon sources. Abbreviations: H4F = tetrahydrofolate; Hyd = group 1f [NiFe]-hydrogenase; pMMO = particulate methane monooxygenase; MDH = methanol dehydrogenase; PQQ = pyrroloquinoline quinone; I = NADH dehydrogenase (complex I); II = succinate dehydrogenase (complex II); IV = cytochrome aa3 oxidase (complex IV). Dashed black lines indicate diffusion. Dashed gray lines indicate unknown process or undetected genes. (c) Molecular model of the putative particulate methane monooxygenase (Pmo) from Candidatus Methylotropicum kingii and comparison with putative catalytic sites of other experimentally validated Pmos. (i) Molecular model of the functional homotrimer of the putative Pmo complex from Ca. M. kingii, shown as a cartoon representation. PmoA, PmoB, and PmoC subunits from one Pmo complex are colored in light blue, dark blue, and sky blue respectively. The other two Pmo complexes in the trimer are colored in transparent yellow. Cu ions bound in the putative active site of the colour-coded Pmo complex are shown as ochre spheres. TM = transmembrane. (ii) Zoomed view of the Pmo complex from panel A, with the three Pmo subunits labelled and the putative active sites CuB and CuC highlighted. (iii) Stick representation of residues in the CuB and CuC active sites of Pmo molecular models from Ca. M. kingii (Gemmatimonadota) and Methylacidiphilum infernorum (Verrucomicrobiota), and from the crystal structure (PDB ID: 3RGB) of Pmo from the Methylococcus capsulatus (Proteobacteria), showing that both sites are conserved between Pmo from Ca. M. kingii and M. capsulatus. Pmo from M. infernorum lacks the CuB site, suggesting this site is not responsible for methane oxidation. (iv) Sequence alignment of metal binding motifs from the CuB (left) and CuC (right) sites from Pmo from Ca. M. kingii (cMk.), M. capsulatus (Mc.), Candidatus Methylomirabilis limnetica (cMl.; Methylomirabilota), and M. infernorum (Mi.). Amino acids involved in Cu coordination are highlighted.

Extended Data Fig. 4 Biogeochemical measurements of in situ soil gas concentrations and validation of static chamber setup.

(a) Mean soil-gas profiles normalized to the respective ambient air concentration (dashed line) of four biological replicates collected at depths of 0, 2, 4, 6, 8, 10, and 16 cm across four Australian ecosystems (n = 112). Note that the different gases were sampled at identical depths, but points are plotted slightly offset on the y-axis for better visibility of error bars. (b) Laboratory static chamber incubations (n = 3) to control for abiotic release of trace gases from chamber’s plastic components. Chamber headspace mixing ratios of hydrogen (H2), carbon monoxide (CO), and methane (CH4) were measured at eight time points during ~25 minutes of incubation time of an inert stainless steel surface. Measurement setup, sampling procedure, and sampling frequency followed those applied during the in situ chamber incubations performed to measured soil-atmosphere flux of trace gases. Circles indicate mixing ratio averaged from three independent incubations, with the vertical lines denoting one standard deviation. Dashed lines indicate the mixing ratio of trace gas averaged across three air samples.

Source data

Extended Data Fig. 5 Maximum-likelihood tree of amino acid sequences of group 3 [NiFe]-hydrogenase large subunits, a marker for hydrogen production during fermentation processes.

The tree shows 129 sequences from soil metagenome-assembled genomes (blue) alongside 172 representative sequences from NCBI reference genomes (black). The subgroup of each reference sequence is denoted according to the HydDB classification scheme40. The tree was constructed using the JTT matrix-based model, used all sites, and was bootstrapped with 50 replicates and midpoint-rooted. All sequences shorter than 350 amino acids were omitted. Source data is provided in Newick (nwk) format.

Extended Data Fig. 6 Gas chromatography studies of one ex situ trace gas consumption experiment.

Shown are four biological replicates collected at depths of 0-5, 5-10, 15-20, and 25-30 cm across four Australian ecosystems (n = 64). Depicted are the oxidation rates recorded over time of (a) atmospheric H2, (b) atmospheric CO, and (c) atmospheric CH4 by soils at each depth compared to heat-killed control soils. Oxidation of the three gases was also measured in the dryland soils following hydration, that is Dryland (wet) samples. Points show average values and vertical lines the error bars representing one standard deviations of four biological replicates. Significance testing of differences in ex situ oxidation rates between ecosystem type and soil depth was carried out using a one-sided Kruskall Wallis H test followed by a pairwise Wilcoxon tests with adjusted p-values using a Benjamini Hochberg correction for multiple testing. Boxplots show median, lower and upper quartile, and minimum and maximum values comparing the ex situ rates of H2, CO, and CH4 oxidation between depths and ecosystems for (d) cell-specific rates and (e) bulk reaction rates per gram of dry soil.

Source data

Extended Data Fig. 7 Copy number of the 16S rRNA gene per gram of soils of four biological replicates.

Soil samples were collected at depths of 0-5, 5-10, 15-20, and 25-30 cm across four Australian ecosystems, with four biological replicates for each depth (n = 64). Boxplots show median, lower and upper quartile, and minimum and maximum values.

Source data

Extended Data Fig. 8 Comparison of the theoretical and measured population of trace gas oxidizers in soils.

Shown are four biological replicates collected at depths of 0-5, 5-10, 15-20, and 25-30 cm across four Australian ecosystems (n = 64). The theoretical populations, in line with the methods of Conrad68, were estimated from thermodynamic modeling and measured ex situ oxidation rates assuming a single maintenance energy for all cells. Two different values were used: 8.9 × 10-15 W cell-1 as assumed by Conrad68 and 1.9 × 10-15 W cell-1 as the median of metabolic rates reported by DeLong et al60. The measured populations of trace gas oxidizers were determined from 16S rRNA gene copy numbers and the relative abundance of the respective functional gene as determined from metagenomics. The ratio of theoretical vs. measured population is equivalent to dividing our calculated power per cell by the respective maintenance energy above. Boxplots show median, lower and upper quartile, and minimum and maximum values.

Source data

Extended Data Fig. 9 Maximum-likelihood tree of amino acid sequences of ribulose 1,5-bisphosphate carboxylase / oxygenase (RbcL), a marker for carbon fixation through the Calvin-Benson cycle.

The tree shows 68 sequences from soil metagenome-assembled genomes (blue) alongside 126 representative sequences from NCBI reference genomes (black). The subtype of each reference sequence is denoted. The tree was constructed using the JTT matrix-based model, used all sites, and was bootstrapped with 50 replicates and midpoint-rooted. Source data is provided in Newick (nwk) format.

Extended Data Fig. 10 Summary of the processes and mediators of trace gas cycling at the soil-atmosphere interface.

The soil bacterial phyla capable of consuming the trace gases H2 (via group 1 and 2 [NiFe]-hydrogenases), CO (via form I carbon monoxide dehydrogenases), and CH4 (via particulate methane monooxygenases) are shown. They are listed in order of the number of metagenome-assembled genomes recovered (black for highly abundant and grey for less abundant trace gas oxidizers). The downward arrows show the net consumption of atmospheric H2, CO, and CH4 by soil bacteria. The upward arrows show processes that result in endogenous production and internal cycling of trace gases. The green boxes indicate biotic processes, whereas the brown boxes indicate abiotic processes.

Supplementary information

Supplementary Information

Supplementary Tables 1–12 and Figs. 1–10.

Reporting Summary

Peer Review Information

Supplementary Tables

Supplementary Tables 1–12.

Supplementary Data

Raw phylogenetic trees in NWK format for Fig. 2a,b, Extended Data Figs. 3, 5 and 9 and Supplementary Figs. 1–10.

Source data

Source Data Fig. 1

Numerical source data.

Source Data Fig. 3

Numerical source data.

Source Data Extended Data Fig. 1

Numerical source data.

Source Data Extended Data Fig. 2

Numerical source data.

Source Data Extended Data Fig. 4

Numerical source data.

Source Data Extended Data Fig. 6

Numerical source data.

Source Data Extended Data Fig. 7

Numerical source data.

Source Data Extended Data Fig. 8

Numerical source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bay, S.K., Dong, X., Bradley, J.A. et al. Trace gas oxidizers are widespread and active members of soil microbial communities. Nat Microbiol 6, 246–256 (2021). https://doi.org/10.1038/s41564-020-00811-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41564-020-00811-w

This article is cited by

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