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:

Microbiomes of microscopic marine invertebrates do not reveal signatures of phylosymbiosis

Abstract

Animals and microorganisms often establish close ecological relationships. However, much of our knowledge about animal microbiomes comes from two deeply studied groups: vertebrates and arthropods. To understand interactions on a broader scale of diversity, we characterized the bacterial microbiomes of close to 1,000 microscopic marine invertebrates from 21 phyla, spanning most of the remaining tree of metazoans. Samples were collected from five temperate and tropical locations covering three marine habitats (sediment, water column and intertidal macroalgae) and bacterial microbiomes were characterized using 16S ribosomal RNA gene sequencing. Our data show that, despite their size, these animals harbour bacterial communities that differ from those in the surrounding environment. Distantly related but coexisting invertebrates tend to share many of the same bacteria, suggesting that guilds of microorganisms preferentially associated with animals, but not tied to any specific host lineage, are the main drivers of the ecological relationship. Host identity is a minor factor shaping these microbiomes, which do not show the same correlation with host phylogeny, or ‘phylosymbiosis’, observed in many large animals. Hence, the current debate on the varying strength of phylosymbiosis within selected lineages should be reframed to account for the possibility that such a pattern might be the exception rather than the rule.

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: Collection of over 1,000 specimens representing most animal phyla.
Fig. 2: Impact of host taxonomy and environmental variables on microbiome composition.
Fig. 3: Microbiome overlap between animals and their environment.
Fig. 4: Microbiome composition differences do not reflect host family identity.
Fig. 5: Weak signals of host genus- and species-specific bacteria.

Similar content being viewed by others

Data availability

All sequence data are deposited in the NCBI Short Read Archive under the BioProject accession number PRJNA746569. Specimen photographs are deposited at Dryad (https://doi.org/10.5061/dryad.ngf1vhhv6).

Code availability

No custom code has been used during this work. All analyses were conducted with publicly accessible packages in R and have been cited in the Methods.

References

  1. Gilbert, S. F., Sapp, J. & Tauber, A. I. A symbiotic view of life: we have never been individuals. Q. Rev. Biol. 87, 325–341 (2012).

    Article  PubMed  Google Scholar 

  2. Bass, D., Stentiford, G. D., Wang, H.-C., Koskella, B. & Tyler, C. R. The pathobiome in animal and plant diseases. Trends Ecol. Evol. 34, 996–1008 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  3. Husnik, F. & Keeling, P. J. The fate of obligate endosymbionts: reduction, integration, or extinction. Curr. Opin. Genet. Dev. 58-59, 1–8 (2019).

    Article  CAS  PubMed  Google Scholar 

  4. Berg, G. et al. Microbiome definition re-visited: old concepts and new challenges. Microbiome 8, 103 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  5. Kwong, W. K. & Moran, N. A. Gut microbial communities of social bees. Nat. Rev. Microbiol. 14, 374–384 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Hammer, T. J., Janzen, D. H., Hallwachs, W., Jaffe, S. P. & Fierer, N. Caterpillars lack a resident gut microbiome. Proc. Nat Acad. Sci. USA 114, 9641–9646 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Holt, C. C., van der Giezen, M., Daniels, C. L., Stentiford, G. D. & Bass, D. Spatial and temporal axes impact ecology of the gut microbiome in juvenile European lobster (Homarus gammarus). ISME J. 14, 531–543 (2020).

    Article  PubMed  Google Scholar 

  8. Pollock, F. J. et al. Coral-associated bacteria demonstrate phylosymbiosis and cophylogeny. Nat. Commun. 9, 4921 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  9. Thomas, T. et al. Diversity, structure and convergent evolution of the global sponge microbiome. Nat. Commun. 7, 11870 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Engelberts, J. P. et al. Characterization of a sponge microbiome using an integrative genome-centric approach. ISME J. 14, 1100–1110 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Ley, R. E. et al. Evolution of mammals and their gut microbes. Science 320, 1647–1651 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Mallot, E. K. & Amato, K. R. Host specificity of the gut microbiome. Nat. Rev. Microbiol. 19, 639–653 (2021).

    Article  CAS  Google Scholar 

  13. Colston, T. J. & Jackson, C. R. Microbiome evolution along divergent branches of the vertebrate tree of life: what is known and unknown. Mol. Ecol. 25, 3776–3800 (2016).

    Article  PubMed  Google Scholar 

  14. Levin, D. et al. Diversity and functional landscapes in the microbiota of animals in the wild. Science 372, eabb5352 (2021).

    Article  CAS  PubMed  Google Scholar 

  15. Nishida, A. H. & Ochman, H. Rates of gut microbiome divergence in mammals. Mol. Ecol. 27, 1884–1897 (2013).

    Article  Google Scholar 

  16. Brooks, A. W., Kohl, K. D., Brucker, R. M., van Opstal, E. J. & Bordenstein, S. R. Phylosymbiosis: relationships and functional effects of microbial communities across host evolutionary history. PLoS Biol. 14, e2000225 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  17. Mazel, F. et al. Is host filtering the main driver of phylosymbiosis across the tree of life? mSystems 3, https://doi.org/10.1128/mSystems.00097-18 (2018).

  18. Lutz, H. L. et al. Ecology and host identity outweigh evolutionary history in shaping the bat microbiome. mBio 4, 6 (2019).

    Google Scholar 

  19. Grond, K. et al. No evidence for phylosymbiosis in Western chipmunk species. FEMS Microbiol. Ecol. 96, fiz182 (2020).

    Article  CAS  PubMed  Google Scholar 

  20. Song, S. J. et al. Comparative analyses of vertebrate gut microbiomes reveal convergence between birds and bats. mBio 11, 1 (2020).

    Article  Google Scholar 

  21. Trevelline, B. K., Sosa, J., Hartup, B. K. & Kohl, K. D. A bird’s-eye view of phylosymbiosis: weak signatures of phylosymbiosis among all 15 species of cranes. Proc. R. Soc. B 287, 20192988 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Muegge, B. D. et al. Diet drives convergence in gut microbiome functions across mammalian phylogeny and within humans. Science 332, 970–974 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Youngblut, N. D. et al. Host diet and evolutionary history explain different aspects of gut microbiome diversity among vertebrate clades. Nat. Commun. 10, 2200 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  24. Amato, K. R. et al. Evolutionary trends in host physiology outweigh dietary niche in structuring primate gut microbiomes. ISME J. 13, 576–587 (2019).

    Article  CAS  PubMed  Google Scholar 

  25. Moeller, A. H. et al. Social behavior shapes the chimpanzee pan-microbiome. Sci. Adv. 2, e1500997 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Eckert, E. M., Anicic, N. & Fontaneto, D. Freshwater zooplankton microbiome composition is highly flexible and strongly influenced by the environment. Mol. Ecol. 30, 1545–1558 (2021).

    Article  PubMed  Google Scholar 

  27. Yatsunenko, T. et al. Human gut microbiome viewed across age and geography. Nature 486, 222–228 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Bik, H. M. Microbial metazoa are microbes too. mSystems 4, e00109–e00119 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  29. Schuelke, T., Pereira, T. J., Hardy, S. M. & Bik, H. M. Nematode-associated microbial taxa do not correlate with host phylogeny, geographic region or feeding morphology in marine sediment habitats. Mol. Ecol. 27, 1930–1951 (2018).

    Article  PubMed  Google Scholar 

  30. Guidetti, R. et al. Further insights in the Tardigrada microbiome: phylogenetic position and prevalence of infection of four new Alphaproteobacteria putative endosymbionts. Zool. J. Linn. Soc. 188, 925–937 (2020).

    Article  Google Scholar 

  31. Giere, O. Meiobenthology (Springer-Verlag, 2009).

  32. Laumer, C. E. et al. Revisiting metazoan phylogeny with genomic sampling of all phyla. Proc. R. Soc. B 286, 20190831 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Hammer, T. J., Sanders, J. G. & Fierer, N. Not all animals need a microbiome. FEMS Microbiol. Lett. 366, fnz117 (2019).

    Article  CAS  PubMed  Google Scholar 

  34. Alejandre-Colomo, C. et al. Cultivable Winogradskyella species are genomically distinct from the sympatric abundant candidate species. ISME Commun. 1, 51 (2021).

    Article  Google Scholar 

  35. Husnik, F. et al. Bacterial and archaeal symbioses with protists. Curr. Biol. 31, R862–R877 (2021).

    Article  CAS  PubMed  Google Scholar 

  36. Salje, J. Cells within cells: Rickettsiales and the obligate intracellular bacterial lifestyle. Nat. Rev. Microbiol. 19, 375–390 (2021).

    Article  CAS  PubMed  Google Scholar 

  37. Neave, M. J., Apprill, A., Ferrier-Pagès, C. & Voolstra, C. R. Diversity and function of prevalent symbiotic marine bacteria in the genus Endozoicomonas. Appl. Microbiol. Biotechnol. 100, 8315–8324 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Weiland-Bräuer, N. et al. Composition of bacterial communities associated with Aurelia aurita changes with compartment, life stage, and population. Appl. Environ. Microbiol. 81, 6038–6052 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  39. Bik, E. M. et al. Marine mammals harbor unique microbiotas shaped by and yet distinct from the sea. Nat. Commun. 7, 10516 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Burns, A. R. et al. Contribution of neutral processes to the assembly of gut microbial communities in the zebrafish over host development. ISME J. 10, 655–664 (2016).

    Article  CAS  PubMed  Google Scholar 

  41. McFall-Ngai, M. Adaptive immunity: care for the community. Nature 445, 153 (2007).

    Article  CAS  PubMed  Google Scholar 

  42. Ruehland, C. & Dubilier, N. Gamma- and epsilonproteobacterial ectosymbionts of a shallow-water marine worm are related to deep-sea hydrothermal vent ectosymbionts. Environ. Microbiol. 12, 2312–2326 (2010).

    CAS  PubMed  Google Scholar 

  43. Gruber-Vodicka, H. R. et al. Two intracellular and cell-type specific bacterial symbionts in the placozoan Trichoplax H2. Nat. Microbiol. 4, 1465–1474 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Schockaert, E. R. in Methods for the Examination of Organismal Diversity in Soils and Sediments (ed. Hall, G. S.) 211–225 (CABI, 1996).

  45. Higgins, R. P. in Introduction to the Study of Meiofauna (eds. Higgins, R. P. and Thiel, H.) 328–331 (SIP, 1988).

  46. Schram, M. D. & Davison, P. G. Irwin Loops—a history and method of constructing homemade loops. Trans. Kans. Acad. Sci. 115, 35–40 (1903).

    Article  Google Scholar 

  47. Medlin, L., Elwood, H. J., Stickel, S. & Sogin, M. L. The characterization of enzymatically amplified eukaryotic 16S-like rRNA-coding regions. Gene 71, 491–499 (1988).

    Article  CAS  PubMed  Google Scholar 

  48. Bower, S. M. et al. Preferential PCR amplification of parasitic protistan small subunit rDNA from metazoan tissues. J. Eukaryot. Microbiol. 51, 325–332 (2004).

    Article  CAS  PubMed  Google Scholar 

  49. Comeau, A. M., Li, W. K. W., Tremblay, J.-E., Carmack, E. C. & Lovejoy, C. Arctic ocean microbial community structure before and after the 2007 record sea ice minimum. PLoS ONE 6, e27492 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Zhang, R.-Y. et al. Design of targeted primers based on 16S rRNA sequences in meta-transcriptomic datasets and identification of a novel taxonomic group in the Asgard archaea. BMC Microbiol. 20, 25 (2020).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  51. Lane, D. J. in Nucleic Acid Techniques in Bacterial Systematics (eds Stackebrandt, E. & Goodfellow, M) 115–175 (Wiley, 1991).

  52. Parada, A. E., Needham, D. M. & Fuhrman, J. A. Every base matters: assessing small subunit rRNA primers for marine microbiomes with mock communities, time series and global field samples. Environ. Microbiol. 18, 1403–1414 (2016).

    Article  CAS  PubMed  Google Scholar 

  53. Marcel, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet J. 17, 10 (2011).

    Article  Google Scholar 

  54. R Core Team. R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, 2020).

  55. Callahan, B. J. DADA2: high-resolution sample inference from Illumina amplicon data. Nat. Methods 13, 581–583 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Wang, Q., Garrity, G. M., Tiedje, J. M. & Cole, J. R. Naïve Bayesian classifier for rapid assignment of rRNA sequences into the new bacterial taxonomy. Appl. Environ. Microbiol. 73, 5261–5267 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Davis, N. M., Proctor, D. M., Holmes, S. P., Relman, D. A. & Callahan, B. J. Simple statistical identification and removal of contaminant sequences in marker-gene and metagenomics data. Microbiome 6, 226 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Wickham, H. ggplot2: Elegant Graphics for Data Analysis (Springer-Verlag, 2016).

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

  61. Kurtz, Z. D. Sparse and compositionally robust inference of microbial ecological networks. PLoS Comput. Biol. 11, e1004226 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  62. Csardi, G. & Nepusz, T. The igraph Software Package for Complex Network Research (InterJournal, 2006).

  63. Breiman, L. Random forests. Mach. Learn. 45, 5–32 (2001).

  64. Kolde, R. pheatmap: pretty heatmaps. R package version 1.0.12 https://CRAN.R-project.org/package=pheatmap (2015).

  65. Lin, H. & Das Peddada, S. Analysis of composition of microbiomes with bias correction. Nat. Commun. 11, 3514 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Oksanen, J. vegan: Community Ecology Package. R package version 2.5.7 https://CRAN.R-project.org/package=vegan (2020).

  67. Rouse, G., Pleijel, F. & Tilic, E. Annelida (Oxford Univ. Press, 2022).

  68. Ahmed, M. & Holovachov, O. Twenty years after De Ley and Blaxter—How far did we progress in understanding the phylogeny of the phylum Nematoda? Animals 11, 3479 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  69. Van Steenkiste, N. W. L., Herbert, E. R. & Leander, B. S. Species diversity in the marine microturbellarian Astrotorhynchus bifidus sensu lato (Platyhelminthes: Rhabdocoela) from the Northeast Pacific Ocean. Mol. Phylogenet. Evol. 120, 259–273 (2018).

Download references

Acknowledgements

We wish to thank F. Bergmeier, B. Brenzinger, S. Fujimoto, U. Jondelius, M. Kolicka, C. Nielsen, A. Schmidt-Rhaesa, M. Sørensen and W. Sterrer for additional taxonomic expertise, the Hakai Institute and CARMABI and their helpful staff (in particular, N. Acharya-Patel, C. Janusson and C. Prentice from Hakai for assistance with DNA extractions) and C. Wall and G. Buckholtz for laboratory procedures at the University of British Columbia. This project was funded by the Tula Foundation’s Hakai Institute (P.J.K. and B.S.L.), the Natural Sciences and Engineering Research Council (NSERC 2019-03896 to B.S.L. and NSERC 2019-03994 to P.J.K.), the Gordon and Betty Moore Foundation (https://doi.org/10.37807/GBMF9201, P.J.K.) and the Canadian Graduate Scholarship programme (N.A.T.I.).

Author information

Authors and Affiliations

Authors

Contributions

The project was conceived and planned by V.B., N.W.L.V.S., M.H., B.S.L. and P.J.K. B.S.L. and P.J.K. are the senior authors and provided supervision and funding. V.B., N.W.L.V.S., M.H. and N.A.T.I. performed most of the field and laboratory work. C.C.H. conducted the analyses. V.M., N.O. and R.S.P. contributed to field and laboratory work. N.W.L.V.S., M.H., P.A.C., K.G., O.H., A.K. and K.W. identified the animals. V.B., C.C.H. and P.J.K. wrote the original draft. All authors edited iterative versions of the final manuscript.

Corresponding authors

Correspondence to V. Boscaro or P. J. Keeling.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Microbiology thanks Diego Fontaneto, Holly Lutz, Owen Wangensteen and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

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

Extended data

Extended Data Fig. 1 Examples of animals collected during the survey.

a, Examples of collected animals from non-arthropod phyla. b, Examples of collected animals from arthropod lineages. Bars stand for 250 μm, except for a7, a9, a11, a12, a13, a15, and a18, where they stand for 50 μm (hatched bars), and a20 and b1, where they stand for 1,000 μm (red bars).

Extended Data Fig. 2 Observed number of ASVs in host-associated and environmental microbial communities.

Bacterial richness, expressed as ASV counts, of the microbial communities, grouped according to: a, host phylum; b, host order within the phylum Annelida and Nematoda; c, location (for both animal-associated and environmental data); and d, habitat (for both animal-associated and environmental data). The dashed box in a separates phyla with less than five sampled specimens each. Microbial community Shannon-diversity (Fig. 2) and ASV numbers are considerably higher in environmental communities than in invertebrate hosts.

Extended Data Fig. 3 Diversity and richness of host-associated and environmental microbial communities do not correlate.

a, Individual animal microbiome Shannon-diversity index and b, ASV counts plotted against the averaged values from corresponding environmental communities. c,d, same plots separated according to host phylum. Differences in Shannon-diversity and richness of animal-associated microbiomes are not tied to Shannon-diversity and richness of background environmental communities. The grey area shows the 95% confidence interval (default geom_smooth se parameter). n = 877 specimens. n of specimens per phylum as in Fig. 2d.

Extended Data Fig. 4 Microbiome overlap between animals and their environment in locations other than Quadra Island.

a-e, Proportion of bacterial Amplicon Sequence Variants shared between individual invertebrates and their environment, separated by habitat for locations with multiple sampled habitats. f-j, Proportion of bacterial ASVs shared between individual specimens and all other animals in the same sample, separated by habitat for locations with multiple sampled habitats. Solid, black lines in circular plots indicate overall average. Dashed lines indicate 25%, 50%, and 75% thresholds, for scale. Black circles plot phylum average.

Extended Data Fig. 5 Influence of keystone environmental bacteria in animal-associated microbiomes isolated from macroalgae and water.

SPIEC-EASI co-occurrence network of key environmental ASVs found in Quadra Island a, macroalgae (n = 253 ASVs) and b, water column (n = 228 ASVs) samples. Each node represents a single ASV. Lines connecting two nodes (edges) indicate an association between the two ASVs. Node size is scaled to eigen-centrality, which considers the number of connecting nodes as well as their subsequent connections. c,d, prevalence and abundance (both as %) of the same environmental ASVs (respective of each habitat) in animals from the same habitat and location. Individual ASVs (on the x axis) are ordered according to their eigen-centrality in the environmental network, and may be represented by multiple datapoints in the abundance plot (on the right) to reflect their varying abundance in multiple host phyla. Grey arrowheads in prevalence plots indicate environmental ASVs that are absent in host-associated microbiomes. Point colour indicates host phyla. As is the case in sediments from the same location (see text), keystone environmental bacteria are not particularly abundant nor prevalent in animal-associated microbiomes.

Extended Data Fig. 6 Correlation of ASVs shared between animals and those shared between animals and the environment.

The proportion of bacterial ASVs shared between individual invertebrates collected in Quadra Island and all other co-occurring animals in the same sample plotted against: a, the proportion of bacterial ASVs shared between animals and their environment; b, the proportion of shared ASVs between animals that are also shared with the environment. Both coloured and separated according to host phylum. While there is a tendency for co-occurring animals to share more ASVs in samples where more ASVs are also shared with the environment, the ASVs responsible for both overlaps do not increase in number accordingly, and hence are not necessarily the same. n of specimens per phylum as in Fig. 2d.

Extended Data Fig. 7 Predicting host taxonomy and environmental factors with random forest models.

Out-of-bag error rates of random forest models using microbial community ASVs to predict potential groupings. From left to right: all data, predicting community type (host-associated vs. environmental); animal-associated data, predicting host phylum, host phylum restricted to phyla that only include more than 20 specimens, host class, and host order; both animal-associated and environmental data, predicting location and habitat. The models can confidently discriminate microbial community type as well as environmental parameters from environmental communities. They fare poorly when discriminating any parameter from host-associated microbiomes, especially those related to host taxonomy.

Extended Data Fig. 8 Phylogenetic relationships used in phylogeny-based phylosymbiosis tests.

Phylogenetic trees (topology only) among a, phyla investigated in this work32. b, families in Annelida;67 c, families in Nematoda;68 d, species in Astrotorhynchus69. Mantel tests, results of which are shown under each panel, compare the phylogenetic topologies shown with similarity dendrograms of microbiomes from corresponding specimens, as shown in d. Microbiome data from specimens belonging to the same taxon (phylum, family, or species) were mapped in 0-branch lengths polytomies in each tree, as shown for Astrotorhynchus (numbers of specimens per polytomy are reported within dark triangles in a–c). All the performed analyses showed a very low degree of covariation (R value) between host phylogeny and microbiome similarity.

Extended Data Fig. 9 Potential phylosymbiosis signal in Echinoderes (Kinorhyncha).

Principal Coordinates Analysis using Bray–Curtis dissimilarity of microbiomes from Echinoderes specimens largely clustered according to host species. Ellipses group specimens of the same species.

Extended Data Fig. 10 Core bacterial families found in invertebrate phyla.

Prevalence of bacterial families at increasing relative abundance thresholds within each invertebrate phylum. Only families present in the majority of specimens (>50%) at or above 0.005% relative abundance are included. Actual prevalence values are included at each threshold with colour denoting degree of prevalence. Families occurring in all specimens at a given abundance threshold (prevalence value = 1) are indicated by a dark grey outline.

Supplementary information

Supplementary Information

Supplementary Notes 1–3 and Figs. 1–3.

Reporting Summary

Supplementary Tables

Supplementary Table 1: Metadata on the 46 samples processed during the survey. Each sample corresponds to 15–29 individual animal specimens and 5–9 environmental aliquots. Supplementary Table 2: Metadata on the 1,000+ invertebrates and 250+ environmental aliquots investigated. Morphology-based taxonomy is only reported for annelids, platyhelminths, gnathostomulids, rotifers, gastrotrichs, bryozoans, kinorhynchs, priapulids, tardigrades, nematodes and hemichordates. Partial 18S ribosomal gene sequences for 232 specimens are also reported. Supplementary Table 3: Analysis of compositions of microbiomes with bias correction (ANCOMBC) estimated separately for each habitat. Log-fold changes with respect to specimens are reported. Standard error, test statistic (standardized effect size), P value (two-way Z-test) and P value adjusted for multiple tests (Holm–Bonferroni method) are also shown for each comparison. Supplementary Table 4: Dataframe listing the prevalence and identity of potential symbionts described in the manuscript. Taxonomic identification indicated by best BLAST hit and DADA2 assignment using the RDP classifier and the SILVA database (v.138). Bootstrap support for taxonomic assignment is also included. Supplementary Table 5: Dataframe listing ASVs assigned to Rickettsiales, Holosporales, Chlamydiae and Endozoicomonas by DADA2 and the RDP classifier using the SILVA database (v.138). Bootstrap support for taxonomic assignment is also included. Best BLAST hit against the full nt database is also reported. Host phylum, location and habitat columns include all occurrences of each ASV.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Boscaro, V., Holt, C.C., Van Steenkiste, N.W.L. et al. Microbiomes of microscopic marine invertebrates do not reveal signatures of phylosymbiosis. Nat Microbiol 7, 810–819 (2022). https://doi.org/10.1038/s41564-022-01125-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41564-022-01125-9

This article is cited by

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