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:

Ligand cross-feeding resolves bacterial vitamin B12 auxotrophies

Abstract

Cobalamin (vitamin B12, herein referred to as B12) is an essential cofactor for most marine prokaryotes and eukaryotes1,2. Synthesized by a limited number of prokaryotes, its scarcity affects microbial interactions and community dynamics2,3,4. Here we show that two bacterial B12 auxotrophs can salvage different B12 building blocks and cooperate to synthesize B12. A Colwellia sp. synthesizes and releases the activated lower ligand α-ribazole, which is used by another B12 auxotroph, a Roseovarius sp., to produce the corrin ring and synthesize B12. Release of B12 by Roseovarius sp. happens only in co-culture with Colwellia sp. and only coincidently with the induction of a prophage encoded in Roseovarius sp. Subsequent growth of Colwellia sp. in these conditions may be due to the provision of B12 by lysed cells of Roseovarius sp. Further evidence is required to support a causative role for prophage induction in the release of B12. These complex microbial interactions of ligand cross-feeding and joint B12 biosynthesis seem to be widespread in marine pelagic ecosystems. In the western and northern tropical Atlantic Ocean, bacteria predicted to be capable of salvaging cobinamide and synthesizing only the activated lower ligand outnumber B12 producers. These findings add new players to our understanding of B12 supply to auxotrophic microorganisms in the ocean and possibly in other ecosystems.

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: Chemical structure of B12 and growth characteristics of Colwellia and Roseovarius when supplemented with B12 or respective B12 building blocks.
Fig. 2: Presence and absence of metabolic pathways related to cobalamin biosynthesis of Colwellia and Roseovarius and related interactions when grown in co-culture.
Fig. 3: Bacterial population dynamics and prophage induction in Roseovarius and Colwellia co-cultures.
Fig. 4: Growth characteristics of C.muelleri grown with Colwellia and/or Roseovarius with respective B12 building blocks.

Similar content being viewed by others

Data availability

Complete genome sequences of Colwellia sp. M166 and Roseovarius sp. M141 were uploaded into the NCBI and annotated by JGI Gold. Genomes can be inquired by their IMG genome identifiers 2828890045 for Colwellia and 2828508468 for Roseovarius. Transcriptomic sequences generated in this study have been deposited into the ENA at EMBL-EBI with the accession number PRJEB43320, using the data brokerage service of the GFBio in compliance with the Minimal Information about any (X) Sequence (MIxS) standard. The data from the sampling stations (ANT-XXVIII/4 and ANT-XXVIII/5) are available from PANGEA under the accession number PANGAEA.906247, and respective sequence data can be retrieved from the ENA under the INSDC accession number PRJEB34453. The SILVA and rfam database was used for transcriptome analysis. For bacterial gene annotation, the ProGenome database was used, and for phage gene and protein annotation, the NR database (from the NCBI), the Virus Orthologous Group database, the InterPro database and the HMM database were used. Source data are provided with this paper.

Code availability

No specialized in-house code was used for this study. All software used for the data analyses in this study are publicly available and cited in the Methods. Custom scripts and the pipeline used have been deposited into GitHub (https://github.com/LeonDlugosch/MetaSeq-Toolkit).

References

  1. Croft, M. T., Lawrence, A. D., Raux-Deery, E., Warren, M. J. & Smith, A. G. Algae acquire vitamin B12 through a symbiotic relationship with bacteria. Nature 438, 90–93 (2005).

    Article  ADS  CAS  PubMed  Google Scholar 

  2. Sañudo-Wilhelmy, S. A., Gómez-Consarnau, L., Suffridge, C. & Webb, E. A. The role of B vitamins in marine biogeochemistry. Annu. Rev. Mar. Sci. 6, 339–367 (2014).

    Article  ADS  Google Scholar 

  3. Shelton, A. N. et al. Uneven distribution of cobamide biosynthesis and dependence in bacteria predicted by comparative genomics. ISME J. 13, 789–804 (2019).

    Article  CAS  PubMed  Google Scholar 

  4. Lu, X., Heal, K. R., Ingalls, A. E., Doxey, A. C. & Neufeld, J. D. Metagenomic and chemical characterization of soil cobalamin production. ISME J. 14, 53–66 (2020).

    Article  CAS  PubMed  Google Scholar 

  5. Menzel, D. W. & Spaeth, J. P. Occurrence of vitamin B12 in the Sargasso Sea. Limnol. Oceanogr. 7, 151–154 (1962).

    Article  ADS  CAS  Google Scholar 

  6. Suffridge, C. P. et al. B vitamins and their congeners as potential drivers of microbial community composition in an oligotrophic marine ecosystem. J. Geophys. Res. Biogeosci. 123, 2890–2907 (2018).

    Article  CAS  Google Scholar 

  7. Cruz-López, R. & Maske, H. The vitamin B1 and B12 required by the marine dinoflagellate Lingulodinium polyedrum can be provided by its associated bacterial community in culture. Front. Microbiol. 7, 560 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  8. Grant, M. A., Kazamia, E., Cicuta, P. & Smith, A. G. Direct exchange of vitamin B12 is demonstrated by modelling the growth dynamics of algal–bacterial cocultures. ISME J. 8, 1418–1427 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Cooper, M. B. et al. Cross-exchange of B-vitamins underpins a mutualistic interaction between Ostreococcus tauri and Dinoroseobacter shibae. ISME J. 13, 334–345 (2019).

    Article  CAS  PubMed  Google Scholar 

  10. Sokolovskaya, O. M., Shelton, A. N. & Taga, M. E. Sharing vitamins: cobamides unveil microbial interactions. Science 369, eaba0165 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Wienhausen, G. et al. Availability of vitamin B12 and its lower ligand intermediate α-ribazole impact prokaryotic and protist communities in oceanic systems. ISME J. 16, 2002–2014 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Sultana, S., Bruns, S., Wilkes, H., Simon, M. & Wienhausen, G. Vitamin B12 is not shared by all marine prototrophic bacteria with their environment. ISME J. 17, 836–845 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Stupperich, E., Steiner, I. & Eisinger, H. J. Substitution of co alpha-(5-hydroxybenzimidazolyl)cobamide (factor III) by vitamin B12 in Methanobacterium thermoautotrophicum. J. Bacteriol. 169, 3076–3081 (1987).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Crofts, T. S., Seth, E. C., Hazra, A. B. & Taga, M. E. Cobamide structure depends on both lower ligand availability and CobT substrate specificity. Chem. Biol. 20, 1265–1274 (2013).

    Article  CAS  PubMed  Google Scholar 

  15. Yi, S. et al. Versatility in corrinoid salvaging and remodeling pathways supports corrinoid-dependent metabolism in Dehalococcoides mccartyi. Appl. Environ. Microbiol. 78, 7745–7752 (2012).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  16. Banerjee, R. & Ragsdale, S. W. The many faces of vitamin B12: catalysis by cobalamin-dependent enzymes. Annu. Rev. Biochem. 72, 209–247 (2003).

    Article  CAS  PubMed  Google Scholar 

  17. Hazra, A. B., Tran, J. L. A., Crofts, T. S. & Taga, M. E. Analysis of substrate specificity in CobT homologs reveals widespread preference for DMB, the lower axial ligand of vitamin B12. Chem. Biol. 20, 1275–1285 (2013).

    Article  CAS  PubMed  Google Scholar 

  18. Helliwell, K. E. et al. Cyanobacteria and eukaryotic algae use different chemical variants of vitamin B12. Curr. Biol. 26, 999–1008 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Heal, K. R. et al. Two distinct pools of B12 analogs reveal community interdependencies in the ocean. Proc. Natl Acad. Sci. USA 114, 364–369 (2017).

    Article  ADS  CAS  PubMed  Google Scholar 

  20. Gray, M. J. & Escalante-Semerena, J. C. A new pathway for the synthesis of α-ribazole-phosphate in Listeria innocua. Mol. Microbiol. 77, 1429–1438 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Gray, M. J. & Escalante-Semerena, J. C. The cobinamide amidohydrolase (cobyric acid-forming) CbiZ enzyme: a critical activity of the cobamide remodeling system of Rhodobacter sphaeroides. Mol. Microbiol. 74, 1198–1210 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Wienhausen, G., Noriega-Ortega, B. E., Niggemann, J., Dittmar, T. & Simon, M. The exometabolome of two model strains of the Roseobacter group: a marketplace of microbial metabolites. Front. Microbiol. 8, 1985 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  23. Johnson, W. M., Kido Soule, M. C. & Kujawinski, E. B. Evidence for quorum sensing and differential metabolite production by a marine bacterium in response to DMSP. ISME J. 10, 2304–2316 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Crofts, T. S., Men, Y., Alvarez-Cohen, L. & Taga, M. E. A bioassay for the detection of benzimidazoles reveals their presence in a range of environmental samples. Front. Microbiol. 5, 592 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  25. Butzin, N. C., Secinaro, M. A., Swithers, K. S., Gogarten, J. P. & Noll, K. M. Thermotoga lettingae can salvage cobinamide to synthesize vitamin B12. Appl. Environ. Microbiol. 79, 7006–7012 (2013).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  26. Woodson, J. D., Zayas, C. L. & Escalante-Semerena, J. C. A new pathway for salvaging the coenzyme B12 precursor cobinamide in archaea requires cobinamide-phosphate synthase (CbiB) enzyme activity. J. Bacteriol. 185, 7193–7201 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Gray, M. J. & Escalante-Semerena, J. C. Single-enzyme conversion of FMNH2 to 5,6-dimethylbenzimidazole, the lower ligand of B12. Proc. Natl Acad. Sci. USA 104, 2921–2926 (2007).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  28. Campbell, G. R. O. et al. Sinorhizobium meliloti bluB is necessary for production of 5,6-dimethylbenzimidazole, the lower ligand of B12. Proc. Natl Acad. Sci. USA 103, 4634–4639 (2006).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  29. Yu, T.-Y. et al. Active site residues critical for flavin binding and 5,6-dimethylbenzimidazole biosynthesis in the flavin destructase enzyme BluB. Protein Sci. 21, 839–849 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Suh, S.-J. & Escalante-Semerena, J. C. Cloning, sequencing and overexpression of cobA which encodes ATP:corrinoid adenosyltranferase in Salmonella typhimurium. Gene 129, 93–97 (1993).

    Article  CAS  PubMed  Google Scholar 

  31. Escalante-Semerena, J. C., Suh, S. J. & Roth, J. R. cobA function is required for both de novo cobalamin biosynthesis and assimilation of exogenous corrinoids in Salmonella typhimurium. J. Bacteriol. 172, 273–280 (1990).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Mireku, S. A. et al. Conformational change of a tryptophan residue in BtuF facilitates binding and transport of cobinamide by the vitamin B12 transporter BtuCD-F. Sci Rep. 7, 41575 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  33. Carini, P. et al. Discovery of a SAR11 growth requirement for thiamin’s pyrimidine precursor and its distribution in the Sargasso Sea. ISME J. 8, 1727–1738 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Wienhausen, G. et al. The overlooked role of a biotin precursor for marine bacteria—desthiobiotin as an escape route for biotin auxotrophy. ISME J. 16, 2599–2609 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Fang, H., Kang, J. & Zhang, D. Microbial production of vitamin B12: a review and future perspectives. Microb. Cell Fact. 16, 15 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  36. Rodionov, D. A., Vitreschak, A. G., Mironov, A. A. & Gelfand, M. S. Comparative genomics of the vitamin B12 metabolism and regulation in prokaryotes. J. Biol. Chem. 278, 41148–41159 (2003).

    Article  CAS  PubMed  Google Scholar 

  37. Kadner, R. J. Repression of synthesis of the vitamin B12 receptor in Escherichia coli. J. Bacteriol. 136, 1050–1057 (1978).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Choi, J., Kotay, S. M. & Goel, R. Various physico-chemical stress factors cause prophage induction in Nitrosospira multiformis 25196—an ammonia oxidizing bacteria. Water Res. 44, 4550–4558 (2010).

    Article  CAS  PubMed  Google Scholar 

  39. Erez, Z. et al. Communication between viruses guides lysis–lysogeny decisions. Nature 541, 488–493 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  40. Silpe, J. E. & Bassler, B. L. A host-produced quorum-sensing autoinducer controls a phage lysis–lysogeny decision. Cell 176, 268–280.e13 (2019).

    Article  CAS  PubMed  Google Scholar 

  41. Laganenka, L. et al. Quorum sensing and metabolic state of the host control lysogeny–lysis switch of bacteriophage T1. mBio 10, e01884–19 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Giovannoni, S. J. et al. Genome streamlining in a cosmopolitan oceanic bacterium. Science 309, 1242–1245 (2005).

    Article  ADS  CAS  PubMed  Google Scholar 

  43. Morris, J. J., Lenski, R. E. & Zinser, E. R. The Black Queen hypothesis: evolution of dependencies through adaptive gene loss. mBio 3, e00036–12 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  44. Bertrand, E. M. et al. Vitamin B12 and iron colimitation of phytoplankton growth in the Ross Sea. Limnol. Oceanogr. 52, 1079–1093 (2007).

    Article  ADS  CAS  Google Scholar 

  45. Bertrand, E. M. et al. Phytoplankton–bacterial interactions mediate micronutrient colimitation at the coastal Antarctic Sea ice edge. Proc. Natl Acad. Sci. USA 112, 9938–9943 (2015).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  46. Tuttle, M. J. & Buchan, A. Lysogeny in the oceans: lessons from cultivated model systems and a reanalysis of its prevalence. Environ. Microbiol. 22, 4919–4933 (2020).

    Article  PubMed  Google Scholar 

  47. Gude, S., Pherribo, G. J. & Taga, M. E. Emergence of metabolite provisioning as a by-product of evolved biological functions. mSystems 5, e00259-20 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  48. Suttle, C. A. Marine viruses—major players in the global ecosystem. Nat. Rev. Microbiol. 5, 801–812 (2007).

    Article  CAS  PubMed  Google Scholar 

  49. Alejandre-Colomo, C., Harder, J., Fuchs, B. M., Rosselló-Móra, R. & Amann, R. High-throughput cultivation of heterotrophic bacteria during a spring phytoplankton bloom in the North Sea. Syst. Appl. Microbiol. 43, 126066 (2020).

    Article  CAS  PubMed  Google Scholar 

  50. Zech, H. et al. Growth phase-dependent global protein and metabolite profiles of Phaeobacter gallaeciensis strain DSM 17395, a member of the marine Roseobacter clade. Proteomics 9, 3677–3697 (2009).

    Article  CAS  PubMed  Google Scholar 

  51. Osterholz, H., Niggemann, J., Giebel, H.-A., Simon, M. & Dittmar, T. Inefficient microbial production of refractory dissolved organic matter in the ocean. Nat. Commun. 6, 7422 (2015).

    Article  ADS  CAS  PubMed  Google Scholar 

  52. Lunau, M., Lemke, A., Dellwig, O. & Simon, M. Physical and biogeochemical controls of microaggregate dynamics in a tidally affected coastal ecosystem. Limnol. Oceanogr. 51, 847–859 (2006).

    Article  ADS  CAS  Google Scholar 

  53. Bakenhus, I. et al. Composition of total and cell-proliferating bacterioplankton community in early summer in the North Sea—Roseobacters are the most active component. Front. Microbiol. 8, 1771 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  54. Fuchs, B. M., Glöckner, F. O., Wulf, J. & Amann, R. Unlabeled helper oligonucleotides increase the in situ accessibility to 16S rRNA of fluorescently labeled oligonucleotide probes. Appl. Environ. Microbiol. 66, 3603–3607 (2000).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  55. Ludwig, W. et al. ARB: a software environment for sequence data. Nucleic Acids Res. 32, 1363–1371 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet.J 17, 10–12 (2011).

    Article  Google Scholar 

  57. Koren, S. et al. Canu: scalable and accurate long-read assembly via adaptive k-mer weighting and repeat separation. Genome Res. 27, 722–736 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Vaser, R., Sović, I., Nagarajan, N. & Šikić, M. Fast and accurate de novo genome assembly from long uncorrected reads. Genome Res. 27, 737–746 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Li, H. Minimap2: pairwise alignment for nucleotide sequences. Bioinformatics 34, 3094–3100 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Karst, S. M., Kirkegaard, R. H. & Albertsen, M. mmgenome: a toolbox for reproducible genome extraction from metagenomes. Preprint at bioRxiv https://doi.org/10.1101/059121 (2016).

  61. R Core Team. R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, 2021).

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

    Article  CAS  PubMed  Google Scholar 

  63. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. 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).

    Article  CAS  PubMed  Google Scholar 

  65. Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  66. Benjamini, Y. & Hochberg, Y. Controlling the false discovery rate: a practical and powerful approach to multiple testing. J. R. Stat. Soc. Ser. B Methodol. 57, 289–300 (1995).

    Article  MathSciNet  Google Scholar 

  67. Markowitz, V. M. et al. IMG: the Integrated Microbial Genomes database and comparative analysis system. Nucleic Acids Res. 40, D115–122 (2012).

    Article  CAS  PubMed  Google Scholar 

  68. Raes, J., Korbel, J. O., Lercher, M. J., von Mering, C. & Bork, P. Prediction of effective genome size in metagenomic samples. Genome Biol. 8, R10 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  69. Brown, C. T. et al. Unusual biology across a group comprising more than 15% of domain Bacteria. Nature 523, 208–211 (2015).

    Article  ADS  CAS  PubMed  Google Scholar 

  70. Dlugosch, L. et al. Significance of gene variants for the functional biogeography of the near-surface Atlantic Ocean microbiome. Nat. Commun. 13, 456 (2022).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  73. Edgar, R. C. Search and clustering orders of magnitude faster than BLAST. Bioinformatics 26, 2460–2461 (2010).

    Article  CAS  PubMed  Google Scholar 

  74. Menzel, P., Ng, K. L. & Krogh, A. Fast and sensitive taxonomic classification for metagenomics with Kaiju. Nat. Commun. 7, 11257 (2016).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  75. Mende, D. R. et al. proGenomes: a resource for consistent functional and taxonomic annotations of prokaryotic genomes. Nucleic Acids Res. 45, D529–D534 (2017).

    Article  CAS  PubMed  Google Scholar 

  76. Brussaard, C. P., Payet, J. P., Winter, C. & Weinbauer, M. G. Quantification of aquatic viruses by flow cytometry. MAVE 11, 102–109 (2010).

    Google Scholar 

  77. Noguchi, H., Taniguchi, T. & Itoh, T. MetaGeneAnnotator: detecting species-specific patterns of ribosomal binding site for precise gene prediction in anonymous prokaryotic and phage genomes. DNA Res. 15, 387–396 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Charif, D. & Lobry, J. R. in Structural Approaches to Sequence Evolution. Biological and Medical Physics, Biomedical Engineering (eds Bastolla, U. et al.) 207–232 (Springer, 2007).

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

    Article  PubMed  PubMed Central  Google Scholar 

  80. Grazziotin, A. L., Koonin, E. V. & Kristensen, D. M. Prokaryotic virus orthologous groups (pVOGs): a resource for comparative genomics and protein family annotation. Nucleic Acids Res. 45, D491–D498 (2017).

    Article  CAS  PubMed  Google Scholar 

  81. Steinegger, M. et al. HH-suite3 for fast remote homology detection and deep protein annotation. BMC Bioinformatics 20, 473 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  82. Kiening, M. et al. Conserved secondary structures in viral mRNAs. Viruses 11, 401 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Finn, R. D. et al. InterPro in 2017—beyond protein family and domain annotations. Nucleic Acids Res. 45, D190–D199 (2017).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Barrero‐Canosa, J., Moraru, C., Zeugner, L., Fuchs, B. M. & Amann, R. Direct-geneFISH: a simplified protocol for the simultaneous detection and quantification of genes and rRNA in microorganisms. Environ. Microbiol. 19, 70–82 (2017).

    Article  PubMed  Google Scholar 

  86. Moraru, C. Gene-PROBER—a tool to design polynucleotide probes for targeting microbial genes. Syst. Appl. Microbiol. 44, 126173 (2021).

    Article  CAS  PubMed  Google Scholar 

  87. Kearse, M. et al. Geneious Basic: an integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 28, 1647–1649 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  88. Barrero-Canosa, J. & Moraru, C. Linking microbes to their genes at single cell level with direct-geneFISH. Methods Mol. Biol. 2246, 169–205 (2021).

    Article  CAS  PubMed  Google Scholar 

  89. Lunau, M., Lemke, A., Walther, K., Martens-Habbena, W. & Simon, M. An improved method for counting bacteria from sediments and turbid environments by epifluorescence microscopy. Environ. Microbiol. 7, 961–968 (2005).

    Article  PubMed  Google Scholar 

  90. Bruns, S., Wienhausen, G., Scholz-Böttcher, B., Heyen, S. & Wilkes, H. Method development and quantification of all B vitamins and selected biosynthetic precursors in winter and spring samples from the North Sea and de novo synthesized by Vibrio campbellii. Mar. Chem. 256, 104300 (2023).

    Article  CAS  Google Scholar 

  91. Bruns, S., Wienhausen, G., Scholz-Böttcher, B. & Wilkes, H. Simultaneous quantification of all B vitamins and selected biosynthetic precursors in seawater and bacteria by means of different mass spectrometric approaches. Anal. Bioanal. Chem. 414, 7839–7854 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Mukherjee, S. et al. Genomes OnLine Database (GOLD) v.8: overview and updates. Nucleic Acids Res. 49, D723–D733 (2021).

    Article  CAS  PubMed  Google Scholar 

  93. Silvester, N. et al. The European Nucleotide Archive in 2017. Nucleic Acids Res. 46, D36–D40 (2018).

    Article  CAS  PubMed  Google Scholar 

  94. Diepenbroek, M. et al. Towards an Integrated Biodiversity and Ecological Research Data Management and Archiving Platform: The German Federation for the Curation of Biological Data (GFBio) (Gesellschaft für Informatik, 2014).

  95. Yilmaz, P. et al. Minimum information about a marker gene sequence (MIMARKS) and minimum information about any (x) sequence (MIxS) specifications. Nat. Biotechnol. 29, 415–420 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We are grateful to C. Alejandre-Colomo, R. Mosseló-Móra, R. Amann and V. Bischoff for providing the North Sea collection of bacterial isolates; B. Kuerzel and M. Wolterink for technical assistance in the HPLC analysis of amino acids and enumerating VLPs; the captains, their crews and the scientific groups of the cruises ANT XXVIII/4 and XXVIII/5 of RV Polarstern and RV Senckenberg for their support; S. Sañudo-Wilhelmy for stimulating discussions regarding B12 analyses; and D. Kirchman for carefully revising a previous version of this publication This work was supported by Deutsche Forschungsgemeinschaft within the Transregional Collaborative Research Center Roseobacter (TRR51) and the Gordon and Betty Moore Foundation (to F.A.).

Author information

Authors and Affiliations

Authors

Contributions

G.W. designed the study, carried out the experiments, the genome analyses for B12 biosynthetic pathways, the transcriptomic analyses and wrote the manuscript. C.M. and G.W. carried out the direct-geneFISH analyses for prophage induction and the prophage-related genomic analyses. D.Q.T. contributed to the CARD-FISH analyses. S.B. and H.W. analysed B12. S.S. assisted in the tripartite consortium experiment. L.D. carried out the transcriptome mapping and mapping of taxa with different B12 genetic traits to the Atlantic Ocean microbiome. F.A. partially supervised the co-culture experiments. M.S. assisted in designing the study, partially supervised the experiments and data analyses and wrote the manuscript, together with G.W. All authors revised the manuscript to finalize it.

Corresponding authors

Correspondence to Gerrit Wienhausen or Meinhard Simon.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature thanks the anonymous reviewers for their contribution to the peer review of this work. Peer reviewer reports are available.

Additional information

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

Extended data figures and tables

Extended Data Fig. 1 Growth characteristics of Colwellia and Roseovarius when supplemented with and without B12 and respective building blocks and methionine and Chaetoceros muelleri in tri-culture with Colwellia and Roseovarius.

a, Maximum growth yield as means of triplicates ± SD as total cell numbers per ml assessed by flow cytometry of Roseovarius; b, of Colwellia when supplemented with B12 (at 1, 10 or 100 pM) or DMB / Cbi (at 1, 10 or 100 pM); Maximum growth yield as means of triplicates ± SD as determined by optical density of c, Roseovarius and d, Colwellia. Each isolate was supplemented with B12 (100 pM), no addition (plain bar) and methionine (1 µM); e, Chlorophyll a fluorescence and total cell numbers enumerated by hemocytometer of axenic Chaetoceros muelleri over time when cultivated without any addition, f, supplemented with B12 at 10 pM and g, 100 pM final concentration and h, when grown in tri-culture with Colwellia or Roseovarius. Experiments a-d were conducted in n = 3 and experiments e-h in n = 4 biological independent samples each in 1 independent experiment.

Source Data

Extended Data Fig. 2 Metabolic pathways related to cobalamin biosynthesis and transcript expression patterns of Colwellia and Roseovarius grown in co-culture.

Differential gene expression between the control, represented by the mono-cultures (Roseovarius & Colwellia) with B12 added, and the co-culture, without B12 addition, was quantified with the DESeq2 package as implemented in R v4.0.5 (R-Core-Team, 2022). In each treatment n = 3 biological independent samples were used. Coloring of genes refers to their LOG-2FC regulation, defined as downregulated (light red to red), upregulated (light green to dark green) and similarly regulated (orange). Genes that are significantly regulated (LOG2-FC larger than 2 and smaller than −2 with an BH adj. p-value < 0.05) are marked with an asterisk. When an illustration represents more than one gene (all genes > (-) 2 LOG-2FC and adj. p-value < 0.05), the asterisk is placed in brackets. Genes that were not identified in the genome (grey). Since the bluB gene in Roseovarius is missing a large section of the sequence, we have indicated this by outlining the gene in gray. Corresponding gene expression significances (adjusted p-value) and further details on genes are presented in Supplementary Data 2. Our documented microbial interactions including the coincidence of prophage induction, host cell lysis and release of B12, whose causative relationships still need to be proven, are illustrated and explained by the dashed arrows and the explanations 1–6 in the legend.

Extended Data Fig. 3 Growth and substrate use of Colwellia and Roseovarius when supplemented with B12 building blocks.

Presented are total bacterial cell numbers assessed by flow cytometry and glutamate (glutamic acid) concentrations of a-d, Colwellia and e-h, Roseovarius supplemented with a, e, B12, b, f, DMB and c, g, Cbi at 1 nM final concentrations each and d, h, without any supplementation and of as well as i, Colwellia + Roseovarius when cultivated in co-culture without any addition. Shown are means of triplicates ± SD. Each condition was carried out in n = 3 biological independent samples and growth characteristics were confirmed in independently conducted experiments.

Source Data

Extended Data Fig. 4 Growth characteristics of Thalassiosira pseudonana and Colwellia and Roseovarius cultivated in consortium.

a, Growth of Colwellia and Roseovarius as monitored by flow cytometric cell counts and calculated by CARD-FISH analyses. Further, numbers of virus-like particles measured by flow cytometry are presented. b, cell numbers of T. pseudonana, assessed microscopically using a hemocytometer, and bacterial cells (black circle) assessed by flow cytometry. c, Total bacterial cell abundance and the ratio of virus-like particle counts to the number of Colwellia. d, Total bacterial cell abundance and the ratio of virus-like particle counts to the number of Roseovarius. e, Relative proportions of Colwellia and Roseovarius cells (% of total abundance) during the culturing time. f. Descriptive illustration of the growth progression of the participating partners of the consortium, T. pseudonana, Roseovarius and Colwellia as well as the phage induction over a period of ten weeks. g, Maximum relative fluorescence of T. pseudonana when cultured with Colwellia and Roseovarius, with addition of 100 pM B12, without any addition, in co-culture with Roseovarius and in co-culture with Colwellia. a-f, Means of triplicates g, and ± SD. Measured values of the respective triplicates can be viewed in the corresponding source data file. Each treatment was carried out in n = 3 biological independent samples and growth characteristics were confirmed in independently conducted tests.

Source Data

Extended Data Fig. 5 Lower ligand building block concentration in the exudate of Colwellia at B12 (20 pM) depleted growth conditions and in North Sea seawater samples.

a, means of triplicates + SD of growth of Colwellia supplemented with 1 nM B12, 20 pM B12 and no addition, monitored by flow cytometric cell counts over time. Extracellular concentrations of α-ribazole and 5,6-dimethylbenzimidazole (DMB) were measured for the growth-limited culture when supplemented with 20 pM B12. b, concentrations of B12, cobinamide, DMB and α -ribazole in North Sea seawater samples as means of triplicates + SD. c, Map of the German Bight of the North Sea. Points mark the sampling stations of the analyzed seawater samples. Experiment a was conducted in n = 3 biological independent samples. Experiment b was conducted on n = 3 independent sampling locations and each sample was detected in n = 3 technical replicates. Experiments a-b were conducted in 1 independent experiment. The map was created using Ocean Data View.

Source Data

Extended Data Fig. 6 Predicted cobamide biosynthesis phenotypes in genomes of marine bacteria and their relative abundance in the Atlantic Ocean.

a, Map showing 22 stations (black dots) from 62°S in the Southern Ocean sector to 47°N in the north Atlantic Ocean sampled at 20 m depth during cruises ANTXXVIII/4 and /5 with RV Polarstern. The percentage given along each pie chart represents the proportion of genome-sequenced prokaryotes of the entire prokaryotic community (identified by 16S rRNA gene). The pie chart of every station sampled presents relative proportions of prokaryotes encoding i) the complete B12 biosynthesis pathway (B12 producers; red), ii) corrin ring biosynthesiser (dark red), iii) cobinamide salvagers (lower ligand producers) (pink) iv) prokaryotes identified as B12 non-producers (grey). Stations are overlayed on a map with annual mean concentrations of chlorophyll a at the surface (https://oceandata.sci.gsfc.nasa.gov). b-e, B12 biosynthesis genes of 1,904 genomes of marine bacteria were analyzed and classified into cobamide biosynthesis phenotypes: as described above for a. Presented are relative abundance of b, all analyzed bacteria, c, phyla and classes, d, orders of Alphaproteobacteria and e, orders of Gammaproteobacteria.

Extended Data Fig. 7 Proportions of genes encoding reactions of B12 biosynthesis in genomes of Rhodobacterales, Alteromonadales, Vibrionales and Betaproteobacteria.

Presence (as percentage) of individual B12 biosynthesis genes in Rhodobacterales (220), Alteromonadales (96), Vibrionales (52) and Betaproteobacteria (126) genomes. Percent of gene abundance is given in color-coded 10% increments from red (0%) to dark green (100%). Genes depicted in grey were not identified in examined genomes.

Extended Data Fig. 8 (Pro)-phage targeting direct-geneFISH signal intensities.

Shown are the mean relative intensities of the (pro)-phage targeting direct-geneFISH signals of Roseovarius in triplicate co-culture samples withdrawn after 96 h and 120 h and the mono-culture Roseovarius samples withdrawn after 96 h as boxplots. Individual, small dots represent the (pro)-phage targeting direct-geneFISH signal intensities of the individually measured regions of interest (ROI; approximate one cell). At least 465 ROIs were defined and validated for each sample. The red dashed line defines the threshold for the definition of induced cells. The percentage below the samples indicates the proportion of induced cells in the total cells analyzed. In each treatment n = 3 biological independent samples were run. The lower and upper hinges of the box correspond to the first and third quartiles (the 25th and 75th percentiles). The central line in the box is the median. The upper whisker extends from the hinge to the largest value no further than 1.5 * IQR from the hinge (where IQR is the inter-quartile range, or distance between the first and third quartiles). The lower whisker extends from the hinge to the smallest value at most 1.5 * IQR of the hinge. Data beyond the end of the whiskers are called “outlying” points and are plotted individually. Prophage induction events, indicated by a significant increase of the phage signal intensity in regions of interest (ROI), were observed in 2 independent experiments.

Source Data

Extended Data Fig. 9 Enumeration of bacterial cells and virus-like particles of Colwellia and Roseovarius grown in co- and mono-culture.

a, Bacterial cell numbers (white triangle) and ratio of virus-like particles to bacterial cell number (black circle) of Colwellia and Roseovarius growing in co-culture, c, Roseovarius and e, Colwellia in mono-culture with B12 supplementation. b, Bacterial cell numbers (white triangle) and virus-like particles (black circle) of Colwellia and Roseovarius growing in co-culture, d, Roseovarius and f, Colwellia in mono-culture with B12 supplementation. Shown are means of triplicates + SD. Each treatment was carried out in n = 3 biological independent samples and growth characteristics were confirmed in independently conducted tests.

Source Data

Supplementary information

Supplementary Figures

Supplementary Figs. 1–5.

Reporting Summary

Peer Review File

Supplementary Table 1

Growth rate, growth yield and required molecules per cell of Colwellia and Roseovarius growing in mono-culture when supplemented with B12 or respective B12 building blocks.

Supplementary Table 2

Intracellular vitamin B12 recovery (Hydroxycobalamin, Cyanocobalamin, Adenosylcobalamin, Methylcobalamin) by LC–MS of Colwellia and Roseovarius growing in mono-culture when supplemented with B12 or respective B12 building blocks.

Supplementary Table 3

Growth rate and growth yield of C. muelleri growing in mono-culture when supplemented with B12 or in consortium with Colwellia and Roseovarius.

Supplementary Table 4

Composition of synASW medium and synASW medium trace elements solution, used for combined diatom–bacteria cultivation of C. muellieri or T.pseudonana with Roseovarius and Colwellia.

Supplementary Table 5

Settings used for selected reaction monitoring.

Supplementary Data 1

Identification of B12 biosynthesis genes, B12-dependent reactions, B12-indipendent reactions, B12 transporter genes and B12 salvage genes in the genomes of Roseovarius and Colwellia.

Supplementary Data 2

Transcriptional gene regulation of both bacterial isolates in co-culture compared with gene regulation when supplemented with the addition of B12 (1 nM) cultivated in mono-culture. Genes transcribed at a log2(FC) below –2 and above 2 are outlined in separate sheets, highlighting respective cellular functions.

Supplementary Data 3

Classification of B12 pathway synthesizing groups of publicly available aquatic bacterial genomes.

Supplementary Data 4

Gene annotation of the prophage Roseophage ICBM167.

Supplementary Data 5

Polynucleotide probe mixture for the detection of Roseophage ICBM167 by targeted direct-geneFISH.

Source data

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

Wienhausen, G., Moraru, C., Bruns, S. et al. Ligand cross-feeding resolves bacterial vitamin B12 auxotrophies. Nature (2024). https://doi.org/10.1038/s41586-024-07396-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1038/s41586-024-07396-y

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: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research