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:

Isolation of an archaeon at the prokaryote–eukaryote interface

Abstract

The origin of eukaryotes remains unclear1,2,3,4. Current data suggest that eukaryotes may have emerged from an archaeal lineage known as ‘Asgard’ archaea5,6. Despite the eukaryote-like genomic features that are found in these archaea, the evolutionary transition from archaea to eukaryotes remains unclear, owing to the lack of cultured representatives and corresponding physiological insights. Here we report the decade-long isolation of an Asgard archaeon related to Lokiarchaeota from deep marine sediment. The archaeon—‘Candidatus Prometheoarchaeum syntrophicum’ strain MK-D1—is an anaerobic, extremely slow-growing, small coccus (around 550 nm in diameter) that degrades amino acids through syntrophy. Although eukaryote-like intracellular complexes have been proposed for Asgard archaea6, the isolate has no visible organelle-like structure. Instead, Ca. P. syntrophicum is morphologically complex and has unique protrusions that are long and often branching. On the basis of the available data obtained from cultivation and genomics, and reasoned interpretations of the existing literature, we propose a hypothetical model for eukaryogenesis, termed the entangle–engulf–endogenize (also known as E3) model.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Fig. 1: Growth curves and photomicrographs of the cultured Lokiarchaeota strain MK-D1.
Fig. 2: Syntrophic amino acid utilization of MK-D1.
Fig. 3: Microscopy characterization and lipid composition of MK-D1.
Fig. 4: Phylogeny of MK-D1 and catabolic features of Asgard archaea.
Fig. 5: Proposed hypothetical model for eukaryogenesis.

Similar content being viewed by others

Data availability

Genomes for Ca. P. syntrophicum MK-D1, Halodesulfovibrio sp. MK-HDV and Methanogenium sp. MK-MG are available under GenBank BioProject accession numbers PRJNA557562, PRJNA557563 and PRJNA557565, respectively. The iTAG sequence data was deposited in BioProject PRJDB8518 with SRA accession numbers DRR184081DRR184101. The 16S rRNA gene sequences of MK-D1, Halodesulfovibrio sp. MK-HDV, Methanogenium sp. MK-MG and clones obtained from primary enrichment culture were deposited in the DDBJ/EMBL/GenBank database under accession numbers LC490619LC490624. The gene expression data of MK-D1 in BioProject PRJDB9032 with the accession number DRR199588. The cryo-electron tomograms of Ca. P. syntrophicum MK-D1 have been deposited in the EMDB with accession codes EMD-0809 and EMD-0852.

References

  1. López-García, P. & Moreira, D. Open questions on the origin of eukaryotes. Trends Ecol. Evol. 30, 697–708 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  2. Martin, W. F., Garg, S. & Zimorski, V. Endosymbiotic theories for eukaryote origin. Phil. Trans. R. Soc. Lond. B 370, 20140330 (2015).

    Article  CAS  Google Scholar 

  3. Eme, L., Spang, A., Lombard, J., Stairs, C. W. & Ettema, T. J. G. Archaea and the origin of eukaryotes. Nat. Rev. Microbiol. 15, 711–723 (2017).

    Article  CAS  PubMed  Google Scholar 

  4. Koonin, E. V. Origin of eukaryotes from within archaea, archaeal eukaryome and bursts of gene gain: eukaryogenesis just made easier? Phil. Trans. R. Soc. Lond. B 370, 20140333 (2015).

    Article  CAS  Google Scholar 

  5. Spang, A. et al. Complex archaea that bridge the gap between prokaryotes and eukaryotes. Nature 521, 173–179 (2015).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  6. Zaremba-Niedzwiedzka, K. et al. Asgard archaea illuminate the origin of eukaryotic cellular complexity. Nature 541, 353–358 (2017).

    Article  ADS  CAS  PubMed  Google Scholar 

  7. Sousa, F. L., Neukirchen, S., Allen, J. F., Lane, N. & Martin, W. F. Lokiarchaeon is hydrogen dependent. Nat. Microbiol. 1, 16034 (2016).

    Article  CAS  PubMed  Google Scholar 

  8. Seitz, K. W., Lazar, C. S., Hinrichs, K.-U., Teske, A. P. & Baker, B. J. Genomic reconstruction of a novel, deeply branched sediment archaeal phylum with pathways for acetogenesis and sulfur reduction. ISME J. 10, 1696–1705 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Dombrowski, N., Teske, A. P. & Baker, B. J. Expansive microbial metabolic versatility and biodiversity in dynamic Guaymas Basin hydrothermal sediments. Nat. Commun. 9, 4999 (2018).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  10. Liu, Y. et al. Comparative genomic inference suggests mixotrophic lifestyle for Thorarchaeota. ISME J. 12, 1021–1031 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Seitz, K. W. et al. Asgard archaea capable of anaerobic hydrocarbon cycling. Nat. Commun. 10, 1822 (2019).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  12. Spang, A. et al. Proposal of the reverse flow model for the origin of the eukaryotic cell based on comparative analyses of Asgard archaeal metabolism. Nat. Microbiol. 4, 1138–1148 (2019).

    Article  CAS  PubMed  Google Scholar 

  13. Pushkarev, A. et al. A distinct abundant group of microbial rhodopsins discovered using functional metagenomics. Nature 558, 595–599 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

  14. Bulzu, P.-A. et al. Casting light on Asgardarchaeota metabolism in a sunlit microoxic niche. Nat. Microbiol. 4, 1129–1137 (2019).

    Article  CAS  PubMed  Google Scholar 

  15. Aoki, M. et al. A long-term cultivation of an anaerobic methane-oxidizing microbial community from deep-sea methane-seep sediment using a continuous-flow bioreactor. PLoS ONE 9, e105356 (2014).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  16. Schink, B. & Stams, A. J. in The Prokaryotes: Prokaryotic Communities and Ecophysiology (eds Rosenberg, E. et al.) 471–493 (Springer, 2013).

  17. Knittel, K., Lösekann, T., Boetius, A., Kort, R. & Amann, R. Diversity and distribution of methanotrophic archaea at cold seeps. Appl. Environ. Microbiol. 71, 467–479 (2005).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  18. Albers, S.-V. & Meyer, B. H. The archaeal cell envelope. Nat. Rev. Microbiol. 9, 414–426 (2011).

    Article  CAS  PubMed  Google Scholar 

  19. Marguet, E. et al. Membrane vesicles, nanopods and/or nanotubes produced by hyperthermophilic archaea of the genus Thermococcus. Biochem. Soc. Trans. 41, 436–442 (2013).

    Article  CAS  PubMed  Google Scholar 

  20. Rosenshine, I., Tchelet, R. & Mevarech, M. The mechanism of DNA transfer in the mating system of an archaebacterium. Science 245, 1387–1389 (1989).

    Article  ADS  CAS  PubMed  Google Scholar 

  21. Imachi, H. et al. Cultivation of methanogenic community from subseafloor sediments using a continuous-flow bioreactor. ISME J. 5, 1913–1925 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Da Cunha, V., Gaia, M., Gadelle, D., Nasir, A. & Forterre, P. Lokiarchaea are close relatives of Euryarchaeota, not bridging the gap between prokaryotes and eukaryotes. PLoS Genet. 13, e1006810 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Da Cunha, V., Gaia, M., Nasir, A. & Forterre, P. Asgard archaea do not close the debate about the universal tree of life topology. PLoS Genet. 14, e1007215 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  24. Spang, A. et al. Asgard archaea are the closest prokaryotic relatives of eukaryotes. PLoS Genet. 14, e1007080 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  25. Brunk, C. F. & Martin, W. F. Archaeal histone contributions to the origin of eukaryotes. Trends Microbiol. 27, 703–714 (2019).

    Article  CAS  PubMed  Google Scholar 

  26. Buckel, W. & Thauer, R. K. Energy conservation via electron bifurcating ferredoxin reduction and proton/Na+ translocating ferredoxin oxidation. Biochim. Biophys. Acta 1827, 94–113 (2013).

    Article  CAS  PubMed  Google Scholar 

  27. Ma, K., Zhou, H. Z. & Adams, M. W. W. Hydrogen production from pyruvate by enzymes purified from the hyperthermophilic archaeon, Pyrococcus furiosus: a key role for NADPH. FEMS Microbiol. Lett. 122, 245–250 (1994).

    Article  CAS  Google Scholar 

  28. Nobu, M. K. et al. The genome of Syntrophorhabdus aromaticivorans strain UI provides new insights for syntrophic aromatic compound metabolism and electron flow. Environ. Microbiol. 17, 4861–4872 (2015).

    Article  CAS  PubMed  Google Scholar 

  29. Martin, W. & Müller, M. The hydrogen hypothesis for the first eukaryote. Nature 392, 37–41 (1998).

    Article  ADS  CAS  PubMed  Google Scholar 

  30. Lyons, T. W., Reinhard, C. T. & Planavsky, N. J. The rise of oxygen in Earth’s early ocean and atmosphere. Nature 506, 307–315 (2014).

    Article  ADS  CAS  PubMed  Google Scholar 

  31. Davín, A. A. et al. Gene transfers can date the tree of life. Nat. Ecol. Evol. 2, 904–909 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  32. Kump, L. R. et al. Isotopic evidence for massive oxidation of organic matter following the great oxidation event. Science 334, 1694–1696 (2011).

    Article  ADS  CAS  PubMed  Google Scholar 

  33. Andersson, S. G. & Kurland, C. G. Origins of mitochondria and hydrogenosomes. Curr. Opin. Microbiol. 2, 535–541 (1999).

    Article  CAS  PubMed  Google Scholar 

  34. Fenchel, T. & Finlay, B. J. Oxygen toxicity, respiration and behavioural responses to oxygen in free-living anaerobic ciliates. J. Gen. Microbiol. 136, 1953–1959 (1990).

    Article  CAS  Google Scholar 

  35. Moreira, D. & López-García, P. Symbiosis between methanogenic archaea and δ-proteobacteria as the origin of eukaryotes: the syntrophic hypothesis. J. Mol. Evol. 47, 517–530 (1998).

    Article  ADS  CAS  PubMed  Google Scholar 

  36. López-García, P. & Moreira, D. Selective forces for the origin of the eukaryotic nucleus. BioEssays 28, 525–533 (2006).

    Article  PubMed  CAS  Google Scholar 

  37. Burns, J. A., Pittis, A. A. & Kim, E. Gene-based predictive models of trophic modes suggest Asgard archaea are not phagocytotic. Nat. Ecol. Evol. 2, 697–704 (2018).

    Article  PubMed  Google Scholar 

  38. Martin, W. F., Tielens, A. G. M., Mentel, M., Garg, S. G. & Gould, S. B. The physiology of phagocytosis in the context of mitochondrial origin. Microbiol. Mol. Biol. Rev. 81, e00008-17 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Baum, D. A. & Baum, B. An inside-out origin for the eukaryotic cell. BMC Biol. 12, 76 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  40. Hutson, S. M. & Rannels, S. L. Characterization of a mitochondrial transport system for branched chain α-keto acids. J. Biol. Chem. 260, 14189–14193 (1985).

    Article  CAS  PubMed  Google Scholar 

  41. Hug, L. A., Stechmann, A. & Roger, A. J. Phylogenetic distributions and histories of proteins involved in anaerobic pyruvate metabolism in eukaryotes. Mol. Biol. Evol. 27, 311–324 (2010).

    Article  CAS  PubMed  Google Scholar 

  42. Degli Esposti, M. et al. Alpha proteobacterial ancestry of the [Fe–Fe]-hydrogenases in anaerobic eukaryotes. Biol. Direct 11, 34 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  43. Pieulle, L. et al. Isolation and characterization of the pyruvate-ferredoxin oxidoreductase from the sulfate-reducing bacterium Desulfovibrio africanus. Biochim. Biophys. Acta 1250, 49–59 (1995).

    Article  PubMed  Google Scholar 

  44. Liebgott, P.-P. et al. Relating diffusion along the substrate tunnel and oxygen sensitivity in hydrogenase. Nat. Chem. Biol. 6, 63–70 (2010).

    Article  CAS  PubMed  Google Scholar 

  45. Winkler, H. H. & Neuhaus, H. E. Non-mitochondrial ATP transport. Trends Biochem. Sci. 24, 64–68 (1999).

    Article  CAS  PubMed  Google Scholar 

  46. Gray, M. W. The pre-endosymbiont hypothesis: a new perspective on the origin and evolution of mitochondria. Cold Spring Harb. Perspect. Biol. 6, a016097 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  47. Villanueva, L., Schouten, S. & Damsté, J. S. S. Phylogenomic analysis of lipid biosynthetic genes of Archaea shed light on the ‘lipid divide’. Environ. Microbiol. 19, 54–69 (2017).

    Article  CAS  PubMed  Google Scholar 

  48. Caforio, A. et al. Converting Escherichia coli into an archaebacterium with a hybrid heterochiral membrane. Proc. Natl Acad. Sci. USA 115, 3704–3709 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Nakamura, K. et al. Application of pseudomurein endoisopeptidase to fluorescence in situ hybridization of methanogens within the family Methanobacteriaceae. Appl. Environ. Microbiol. 72, 6907–6913 (2006).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  50. Cevc, G. & Richardsen, H. Lipid vesicles and membrane fusion. Adv. Drug Deliv. Rev. 38, 207–232 (1999).

    Article  CAS  PubMed  Google Scholar 

  51. Nunoura, T. et al. Microbial diversity in deep-sea methane seep sediments presented by SSU rRNA gene tag sequencing. Microbes Environ. 27, 382–390 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  52. Toki, T., Higa, R., Ijiri, A., Tsunogai, U. & Ashi, J. Origin and transport of pore fluids in the Nankai accretionary prism inferred from chemical and isotopic compositions of pore water at cold seep sites off Kumano. Earth Planets Space 66, 137 (2014).

    Article  ADS  Google Scholar 

  53. Nakahara, N. et al. Aggregatilinea lenta gen. nov., sp. nov., a slow-growing, facultatively anaerobic bacterium isolated from subseafloor sediment, and proposal of the new order Aggregatilineales ord. nov. within the class Anaerolineae of the phylum Chloroflexi. Int. J. Syst. Evol. Microbiol. 69, 1185–1194 (2019).

    Article  CAS  PubMed  Google Scholar 

  54. Murakami, S., Fujishima, K., Tomita, M. & Kanai, A. Metatranscriptomic analysis of microbes in an oceanfront deep-subsurface hot spring reveals novel small RNAs and type-specific tRNA degradation. Appl. Environ. Microbiol. 78, 1015–1022 (2012).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  55. Imachi, H. et al. Cultivable microbial community in 2-km-deep, 20-million-year-old subseafloor coalbeds through ~1000 days anaerobic bioreactor cultivation. Sci. Rep. 9, 2305 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  56. Miyashita, A. et al. Development of 16S rRNA gene-targeted primers for detection of archaeal anaerobic methanotrophs (ANMEs). FEMS Microbiol. Lett. 297, 31–37 (2009).

    Article  CAS  PubMed  Google Scholar 

  57. Yamaguchi, T. et al. In situ DNA-hybridization chain reaction (HCR): a facilitated in situ HCR system for the detection of environmental microorganisms. Environ. Microbiol. 17, 2532–2541 (2015).

    Article  CAS  PubMed  Google Scholar 

  58. Miyazaki, M. et al. Sphaerochaeta multiformis sp. nov., an anaerobic, psychrophilic bacterium isolated from subseafloor sediment, and emended description of the genus Sphaerochaeta. Int. J. Syst. Evol. Microbiol. 64, 4147–4154 (2014).

    Article  PubMed  CAS  Google Scholar 

  59. Toyooka, K. et al. Wide-range high-resolution transmission electron microscopy reveals morphological and distributional changes of endomembrane compartments during log to stationary transition of growth phase in tobacco BY-2 cells. Plant Cell Physiol. 55, 1544–1555 (2014).

    Article  CAS  PubMed  Google Scholar 

  60. Kremer, J. R., Mastronarde, D. N. & McIntosh, J. R. Computer visualization of three-dimensional image data using IMOD. J. Struct. Biol. 116, 71–76 (1996).

    Article  CAS  PubMed  Google Scholar 

  61. Takano, Y. et al. Insight into anaerobic methanotrophy from 13C/12C- amino acids and 14C/12C-ANME cells in seafloor microbial ecology. Sci. Rep. 8, 14070 (2018).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  62. Okumura, T. et al. Hydrogen and carbon isotope systematics in hydrogenotrophic methanogenesis under H2-limited and H2-enriched conditions: implications for the origin of methane and its isotopic diagnosis. Prog. Earth Planet. Sci. 3, 14 (2016).

    Article  ADS  Google Scholar 

  63. Takano, Y., Kashiyama, Y., Ogawa, N. O., Chikaraishi, Y. & Ohkouchi, N. Isolation and desalting with cation-exchange chromatography for compound-specific nitrogen isotope analysis of amino acids: application to biogeochemical samples. Rapid Commun. Mass Spectrom. 24, 2317–2323 (2010).

    Article  ADS  CAS  PubMed  Google Scholar 

  64. Chikaraishi, Y. et al. Instrumental Optimization for Compound-specific Nitrogen Isotope Analysis of Amino Acids by Gas Chromatography/Combustion/Isotope Ratio Mass Spectrometry in Earth, Life and Isotopes (eds Ohkouchi, N. et al.) 367–386 (Kyoto Univ. Press, 2010).

  65. Leggett, R. M., Clavijo, B. J., Clissold, L., Clark, M. D. & Caccamo, M. NextClip: an analysis and read preparation tool for Nextera long mate pair libraries. Bioinformatics 30, 566–568 (2014).

    Article  CAS  PubMed  Google Scholar 

  66. Bankevich, A. et al. SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. J. Comput. Biol. 19, 455–477 (2012).

    Article  MathSciNet  CAS  PubMed  PubMed Central  Google Scholar 

  67. Lin, H.-H. & Liao, Y.-C. Accurate binning of metagenomic contigs via automated clustering sequences using information of genomic signatures and marker genes. Sci. Rep. 6, 24175 (2016).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  68. Boetzer, M., Henkel, C. V., Jansen, H. J., Butler, D. & Pirovano, W. Scaffolding pre-assembled contigs using SSPACE. Bioinformatics 27, 578–579 (2011).

    Article  CAS  PubMed  Google Scholar 

  69. Seemann, T. Prokka: rapid prokaryotic genome annotation. Bioinformatics 30, 2068–2069 (2014).

    Article  CAS  PubMed  Google Scholar 

  70. Marchler-Bauer, A. & Bryant, S. H. CD-Search: protein domain annotations on the fly. Nucleic Acids Res. 32, W327–W331 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Marchler-Bauer, A. et al. CDD: NCBI’s conserved domain database. Nucleic Acids Res. 43, D222–D226 (2015).

    Article  CAS  PubMed  Google Scholar 

  72. Jones, P. et al. InterProScan 5: genome-scale protein function classification. Bioinformatics 30, 1236–1240 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Petersen, T. N., Brunak, S., von Heijne, G. & Nielsen, H. SignalP 4.0: discriminating signal peptides from transmembrane regions. Nat. Methods 8, 785–786 (2011).

    Article  CAS  PubMed  Google Scholar 

  74. Yin, Y. et al. dbCAN: a web resource for automated carbohydrate-active enzyme annotation. Nucleic Acids Res. 40, W445–W451 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Rawlings, N. D., Barrett, A. J. & Finn, R. Twenty years of the MEROPS database of proteolytic enzymes, their substrates and inhibitors. Nucleic Acids Res. 44, D343–D350 (2016).

    Article  CAS  PubMed  Google Scholar 

  76. Fischer, M. & Pleiss, J. The Lipase Engineering Database: a navigation and analysis tool for protein families. Nucleic Acids Res. 31, 319–321 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Søndergaard, D., Pedersen, C. N. S. & Greening, C. HydDB: a web tool for hydrogenase classification and analysis. Sci. Rep. 6, 34212 (2016).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  78. Boutet, E., Lieberherr, D., Tognolli, M., Schneider, M. & Bairoch, A. UniProtKB/Swiss-Prot. Methods Mol. Biol. 406, 89–112 (2007).

    CAS  PubMed  Google Scholar 

  79. Lima, T. et al. HAMAP: a database of completely sequenced microbial proteome sets and manually curated microbial protein families in UniProtKB/Swiss-Prot. Nucleic Acids Res. 37, D471–D478 (2009).

    Article  CAS  PubMed  Google Scholar 

  80. Katoh, K. & Standley, D. M. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol. Biol. Evol. 30, 772–780 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Kozlov, A. M., Darriba, D., Flouri, T., Morel, B. & Stamatakis, A. RAxML-NG: a fast, scalable and user-friendly tool for maximum likelihood phylogenetic inference. Bioinformatics 35, 4453–4455 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Ronquist, F. et al. MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model space. Syst. Biol. 61, 539–542 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  83. Pruesse, E., Peplies, J. & Glöckner, F. O. SINA: accurate high-throughput multiple sequence alignment of ribosomal RNA genes. Bioinformatics 28, 1823–1829 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Quast, C. et al. The SILVA ribosomal RNA gene database project: improved data processing and web-based tools. Nucleic Acids Res. 41, D590–D596 (2013).

    Article  CAS  PubMed  Google Scholar 

  85. 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 

  86. Camacho, C. et al. BLAST+: architecture and applications. BMC Bioinformatics 10, 421 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. UniProt Consortium. UniProt: a worldwide hub of protein knowledge. Nucleic Acids Res. 47, D506–D515 (2019).

    Article  CAS  Google Scholar 

  89. Edgar, R. C. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 32, 1792–1797 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Capella-Gutiérrez, S., Silla-Martínez, J. M. & Gabaldón, T. trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25, 1972–1973 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  91. Guindon, S. et al. New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0. Syst. Biol. 59, 307–321 (2010).

    Article  CAS  PubMed  Google Scholar 

  92. Price, M. N., Dehal, P. S. & Arkin, A. P. FastTree 2—approximately maximum-likelihood trees for large alignments. PLoS ONE 5, e9490 (2010).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  93. Bolger, A. M., Lohse, M. & Usadel, B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics 30, 2114–2120 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Magoc, T., Wood, D. & Salzberg, S. L. EDGE-pro: estimated degree of gene expression in prokaryotic genomes. Evol. Bioinform. Online 9, 127–136 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  95. Axley, M. J. & Grahame, D. A. Kinetics for formate dehydrogenase of Escherichia coli formate-hydrogenlyase. J. Biol. Chem. 266, 13731–13736 (1991).

    Article  CAS  PubMed  Google Scholar 

  96. Itoh, T., Suzuki, K. & Nakase, T. Thermocladium modestius gen. nov., sp. nov., a new genus of rod-shaped, extremely thermophilic crenarchaeote. Int. J. Syst. Bacteriol. 48, 879–887 (1998).

    Article  PubMed  Google Scholar 

  97. Zillig, W. et al. The archaebacterium Thermofilum pendens represents, a novel genus of the thermophilic, anaerobic sulfur respiring Thermoproteales. Syst. Appl. Microbiol. 4, 79–87 (1983).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank H. Ohno and T. Yamaguchi for assistance with HCR-FISH analysis; T. Terada for help with NanoSIMS sample preparation; M. Isozaki for assistance with cultivation experiments; T. Kubota for assistance with chemical analysis; K. Takishita, A. Yabuki, T. Shiratori, A. Ohashi, F. Inagaki, T. Nunoura, S. Kawagucci, T. Shibuya, S. Ishii, S. Suzuki, Y. Tsukatani, C. Chen, Y. Kuruma and R. C. Robinson for advice and discussion; A. Miyashita, Y. Yashiro, K. Aoi, M. Ehara, M. Aoki and Y. Saito for assistance with operating the bioreactor; and J. Ashi and the RV Yokosuka and RV Shinkai 6500 operation team during cruise YK06-03 (JAMSTEC) and the shipboard scientists and crews of the RV Chikyu Shakedown Cruise CK06-06 for their assistance in collecting samples. This study was partially supported by grants from the Japan Society for the Promotion of Science (JSPS) (KAKENHI grants 18687006, 21687006, 24687011, 15H02419 and 19H01005 to H.I., 18H03367 to M.K.N., 26710012, 18H02426, 18H05295 to H.T., 18H04468 and 18K18795 to M.I. and Grant-in-Aid for JSPS Fellow 16J10845 to N.N.). This work was also supported by JSPS KAKENHI grant number JP16H06280, Grant-in-Aid for Scientific Research on Innovative Areas–Platforms for Advanced Technologies and Research Resources ‘Advanced Bioimaging Support’ and the Cooperative Study Program (19-504) of National Institute for Physiological Sciences.

Author information

Authors and Affiliations

Authors

Contributions

H.I. conceived the study and carried out the deep-marine sediment sampling. H.I., N.N., M.O., M.M. and S.S. conducted cultivation and culture-based experiments. M.K.N. performed metabolic reconstruction and phylogenetic analyses. M.K.N. and Y. Takaki performed genome analysis. H.I., N.N., Y. Morono, M.O., T.I., M.I., K.M., C.S. and K.U. carried out the microscopy and NanoSIMS work. M.O., Y.S. and Y.Y. performed qPCR, SSU rRNA gene analysis and DNA/RNA sequencing. Y. Takano, Y. Matsui and E.T. performed chemical analysis. H.I., M.K.N., N.N., Y. Morono, Y. Takaki, Y. Takano, K.M., C.S., T.Y., Y.K., H.T. and K.T. conducted data interpretation. H.I., M.K.N., Y. Takano, H.T., Y.K. and K.T. wrote the manuscript with input from all co-authors. All authors have read and approved the manuscript submission.

Corresponding authors

Correspondence to Hiroyuki Imachi or Masaru K. Nobu.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature thanks Sonja-Verena Albers, Petr G. Leiman, James McInerney, Christa Schleper and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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 Growth of MK-D1.

a, Effect of temperature on growth of MK-D1. Data are mean ± s.d. of triplicate determinations. Each data point is shown as a dot. The temperature range test was performed twice with similar results. b, c, The amino acid concentrations and growth curves of MK-D1 in pure cocultures at 20 °C. Results from cultures 1 (b) and 2 (c) are shown. Please note that the initial concentrations of amino acids were normalized to 100%. Total amino acids and several representative amino acids (Val, valine; Leu, leucine; Ile, isoleucine) are independently shown for the duplicate culture samples. Detailed iTAG-based community compositions of the cultures are shown in Supplementary Table 1.

Extended Data Fig. 2 Circular representation of MK-D1 genome.

From the outside to the centre: the distribution of the coding sequences based on the conserved (orange) or non-conserved (grey) genes in the first circle, non-coding RNAs in the second circle, GC content showing deviation from average (40.7%) in the third circle, and GC skew in the fourth circle. The GC content and GC skew were calculated using a sliding window of 2 kb in step of 10 kb. The coding sequences and RNA genes illustrate the findings for plus and minus strands.

Extended Data Fig. 3 Other representative photomicrographs of MK-D1 cultures and Methanobacterium sp. strain MO-MB1.

a, b, Fluorescence images of cells from enrichment cultures after 8 (a) and 11 (b) transfers stained with DAPI (violet) and hybridized with nucleotide probes that target MK-D1 (green) and Bacteria (red). The images are different fields of view to those shown in Fig. 1b, c, which were taken at the same time. c, A fluorescence image of cells in the enrichments after 11 transfers hybridized with nucleotide probes that target MK-D1 (green) and Archaea (but with one mismatch against MK-D1; red). Large and irregular coccoid-shaped cells stained by only ARC915 are probably Methanogenium. d, e, Dividing cells of MK-D1 with a bleb. The top-right inset image in e shows a magnification of the bleb. f, g, Cryo-EM images of MK-D1 cells and large membrane vesicles (white arrows). h, i, Ultrathin sections of MK-D1 cells with a membrane vesicle. The image i shows a magnified image of h. j, k, SEM images of MK-D1 cells with protrusions. l, Ultrathin section of a MK-D1 cell with a protrusion. m, n, Photomicrographs of pure culture of Methanobacterium sp. strain MO-MB1 cells stained with SYBR Green I. Phase-contrast (m) and fluorescence (n) images of the same field are shown. a, b, The FISH experiments were performed three times with similar results. d, e, j, k, The SEM images are representative of n = 122 recorded images that were obtained from four independent observations from four culture samples. The lipid composition experiments were repeated twice and gave similar results. f, g, The cryo-EM images are representative of n = 14 recorded images that were taken from two independent observations from two culture samples. h, i, l, The ultrathin-section images are representative of n = 131 recorded images that were obtained from six independent observations from six culture samples. m, n, The SYBR Green I staining experiment was performed once, but all 10 recorded images showed similar results. Detailed iTAG analyses of cultures are shown in Supplementary Table 1.

Extended Data Fig. 4 Ribosomal protein- and 16S rRNA gene-based phylogeny of MK-D1.

a, Phylogenomic tree of MK-D1 and select cultured archaea, eukaryotes and bacteria based on 31 ribosomal proteins conserved across the three domains (Supplementary Table 7). Ribosomal protein sequences of MK-D1, the organisms shown in the tree and MAGs of uncultured archaeal lineages (Supplementary Table 8) were aligned individually using MAFFT. MAG-derived sequences (except for Ca. Korarchaeum) were then removed for tree construction. After removing all-gap positions and concatenation, the maximum-likelihood tree was constructed using RAxML-NG. Bootstrap values around critical branching points are also shown. In total, 14,875 sites of the alignment were used for tree construction. b, A ribosomal protein-based phylogenomic tree constructed using MrBayes. Bayesian inference phylogenies were calculated using MrBayes 3.2.7a and a ribosomal protein concatenated alignment used for Fig. 4a. c, Phylogenetic tree of MK-D1 and related archaea based on 16S rRNA genes. The 16S rRNA gene sequences were aligned using SINA against the Silva v.132 alignment and the maximum-likelihood tree was calculated using RAxML.

Extended Data Fig. 5 Amino acid, cofactor and nucleotide biosynthesis capacities of MK-D1 and other Asgard archaea.

Genomes that encode proteins for the synthesis of amino acids, cofactors and nucleotides from pyruvate or acetyl-CoA (dark blue) and synthesis from other intermediates (light blue) are indicated. Those without complete pathways from pyruvate and/or acetyl-CoA are indicated in white. Halodesulfovibrio sp. strain MK-HDV and Methanogenium sp. strain MK-MG isolated in this study are also shown.

Extended Data Fig. 6 Maximum-likelihood tree of Asgard archaea urocanate hydratase.

Urocanate hydratase (HutU) homologues were obtained by BLASTp analysis of the Asgard archaea sequences against the UniProt database (release 2019_06). Of homologues with sequence similarity ≥40% and overlap ≥70%, representative sequences were selected using CD-HIT with a clustering cut-off of 70% similarity (otherwise default settings were used). Additional homologues with verified biochemical activity, sequence similarity ≥30% and overlap ≥70% were obtained by BLASTp analysis of the Asgard archaea sequences against the UniProt/SwissProt database (2019_05). Sequences were aligned using MAFFT v.7 with default settings and trimmed using trimAl v.1.2 with default settings. The maximum-likelihood tree was constructed using RAxML-NG using fixed empirical substitution matrix (LG), 4 discrete GAMMA categories, empirical amino acid frequencies from the alignment and 100 bootstrap replicates. In total, 876 sites of the alignment were used for tree construction.

Extended Data Fig. 7 Maximum-likelihood tree of Asgard archaea l-threonine/l-serine dehydratase.

a, Tree calculated for target Asgard archaea l-threonine/l-serine dehydratase (TdcB) and homologues. TdcB homologues were obtained by BLASTp analysis of the Asgard archaea sequences against the UniProt reference proteome and SwissProt database (release 2019_06). Of homologues with sequence similarity ≥40%, overlap ≥70% and predicted prosite domain PS00165 (serine/threonine dehydratases pyridoxal-phosphate attachment site), representative sequences were selected using CD-HIT with a clustering cut-off of 70% similarity (otherwise default settings were used). Additional homologues with verified biochemical activity, sequence similarity ≥30% and overlap ≥70% were obtained by BLASTp analysis of the Asgard archaea sequences against the UniProt/SwissProt database (2019_05). Sequences were aligned using MAFFT v.7 with default settings. Positions with gaps in more than 10% of the sequences were excluded from the alignment using trimAl v.1.2 (-gt 0.9; and otherwise default settings were used). The maximum-likelihood tree was constructed using PhyML using a fixed empirical substitution matrix (LG), 4 discrete GAMMA categories, empirical amino acid frequencies from the alignment and 100 bootstrap replicates (-b 100 -d aa -m LG -v e). In total, 308 sites of the alignment were used for tree construction. b, Tree calculated for a subset of sequences contained in a section of the original tree (branches that are coloured blue). Sequences were realigned and trimmed as described for a. In total, 308 sites of the alignment were used for tree construction.

Extended Data Table 1 SSU rRNA gene clones obtained from the primary and six successive transferred enrichment cultures
Extended Data Table 2 Carbon isotope fractionation values in MK-D1 cultures after 120 days incubation with and without stable isotope labelled amino acids
Extended Data Table 3 Growth of MK-D1 after incubation of 120 days with a range of substrates

Supplementary information

Supplementary Information

This file contains Supplementary Notes 1–9, Supplementary Methods, Supplementary Figures 1–18, and Supplementary References.

Reporting Summary

Supplementary Tables

This file contains Supplementary Tables 1–10

Supplementary Video 1

| Tilt-series images of a single cell of MK-D1

Supplementary Video 2

| Z-slices of the tomographic three-dimensional reconstruction from the tilt-series in Supplementary Video 1

Supplementary Video 3

| Animation of the same MK-D1 cell as in Supplementary Video 2 The cell envelope and membrane vesicles are colored in light blue and pink, respectively.

Supplementary Video 4

| Tilt-series images of MK-D1 cells

Supplementary Video 5

| Z-slices of the tomographic three-dimensional reconstruction from the tilt-series in Supplementary Video 4

Supplementary Video 6

| Animation of the same MK-D1 cells as in Supplementary Video 5 The cell envelope and membrane vesicles are colored in light blue and pink, respectively.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Imachi, H., Nobu, M.K., Nakahara, N. et al. Isolation of an archaeon at the prokaryote–eukaryote interface. Nature 577, 519–525 (2020). https://doi.org/10.1038/s41586-019-1916-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-019-1916-6

This article is cited by

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

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

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