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:

Bacteroides fragilis ubiquitin homologue drives intraspecies bacterial competition in the gut microbiome

Abstract

Interbacterial antagonism and associated defensive strategies are both essential during bacterial competition. The human gut symbiont Bacteroides fragilis secretes a ubiquitin homologue (BfUbb) that is toxic to a subset of B. fragilis strains in vitro. In the present study, we demonstrate that BfUbb lyses certain B. fragilis strains by non-covalently binding and inactivating an essential peptidyl-prolyl isomerase (PPIase). BfUbb-sensitivity profiling of B. fragilis strains revealed a key tyrosine residue (Tyr119) in the PPIase and strains that encode a glutamic acid residue at Tyr119 are resistant to BfUbb. Crystal structural analysis and functional studies of BfUbb and the BfUbb–PPIase complex uncover a unique disulfide bond at the carboxy terminus of BfUbb to mediate the interaction with Tyr119 of the PPIase. In vitro coculture assays and mouse studies show that BfUbb confers a competitive advantage for encoding strains and this is further supported by human gut metagenome analyses. Our findings reveal a previously undescribed mechanism of bacterial intraspecies competition.

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: BfUbb exhibits potent bactericidal toxicity depending on the C-terminal disulfide bond.
Fig. 2: BfUbb functions in the periplasm causing elongation, swelling and lysis of sensitive strains.
Fig. 3: BfUbb targets a periplasmic PPIase to antagonize sensitive B. fragilis.
Fig. 4: Determination of BfUbb–PPIase binding specificity.
Fig. 5: BfUbb exerts bactericidal toxicity by impairing the prolyl isomerase and chaperone activities of targeted essential PPIase.
Fig. 6: BfUbb increases the competitiveness of its encoding strains in intraspecies competition.

Similar content being viewed by others

Data availability

The atomic coordinates and structure factors generated in the present study have been deposited in the PDB under the accession nos. 8HM1, 8HM2, 8HM3 and 8HM4. The whole-genome sequence of B. fragilis GS077 and GS084 was deposited into GenBank under accession nos. JAVFHL000000000 (for GS077) and CP133097 (for GS084). Source data are provided with this paper.

References

  1. Faith, J. J. et al. The long-term stability of the human gut microbiota. Science 341, 1237439 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  2. Peterson, S. B., Bertolli, S. K. & Mougous, J. D. The central role of interbacterial antagonism in bacterial life. Curr. Biol. 30, R1203–r1214 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. García-Bayona, L. & Comstock, L. E. Bacterial antagonism in host-associated microbial communities. Science https://doi.org/10.1126/science.aat2456 (2018).

  4. Wexler, A. G. & Goodman, A. L. An insider’s perspective: Bacteroides as a window into the microbiome. Nat. Microbiol. 2, 17026 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Chatzidaki-Livanis, M. et al. Gut symbiont Bacteroides fragilis secretes a eukaryotic-like ubiquitin protein that mediates intraspecies antagonism. mBio https://doi.org/10.1128/mBio.01902-17 (2017).

  6. Shumaker, A. M., Laclare McEneany, V., Coyne, M. J., Silver, P. A. & Comstock, L. E. Identification of a fifth antibacterial toxin produced by a single Bacteroides fragilis strain. J. Bacteriol. https://doi.org/10.1128/jb.00577-18 (2019).

  7. McEneany, V. L., Coyne, M. J., Chatzidaki-Livanis, M. & Comstock, L. E. Acquisition of MACPF domain-encoding genes is the main contributor to LPS glycan diversity in gut Bacteroides species. ISME J. 12, 2919–2928 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Roelofs, K. G., Coyne, M. J., Gentyala, R. R., Chatzidaki-Livanis, M. & Comstock, L. E. Bacteroidales secreted antimicrobial proteins target surface molecules necessary for gut colonization and mediate competition in vivo. mBio https://doi.org/10.1128/mBio.01055-16 (2016).

  9. Chatzidaki-Livanis, M., Coyne, M. J. & Comstock, L. E. An antimicrobial protein of the gut symbiont Bacteroides fragilis with a MACPF domain of host immune proteins. Mol. Microbiol. 94, 1361–1374 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Bao, Y. et al. A common pathway for activation of host-targeting and bacteria-targeting toxins in human intestinal bacteria. mBio 12, e0065621 (2021).

    Article  PubMed  Google Scholar 

  11. Evans, J. C. et al. A proteolytically activated antimicrobial toxin encoded on a mobile plasmid of Bacteroidales induces a protective response. Nat. Commun. 13, 4258 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Matano, L. M., Coyne, M. J., García-Bayona, L. & Comstock, L. E. Bacteroidetocins target the essential outer membrane protein BamA of Bacteroidales symbionts and pathogens. mBio 12, e0228521 (2021).

    Article  PubMed  Google Scholar 

  13. Patrick, S. et al. A unique homologue of the eukaryotic protein-modifier ubiquitin present in the bacterium Bacteroides fragilis, a predominant resident of the human gastrointestinal tract. Microbiology 157, 3071–3078 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Bitto, E. & McKay, D. B. Crystallographic structure of SurA, a molecular chaperone that facilitates folding of outer membrane porins. Structure 10, 1489–1498 (2002).

    Article  CAS  PubMed  Google Scholar 

  15. Calabrese, A. N. et al. Inter-domain dynamics in the chaperone SurA and multi-site binding to its outer membrane protein clients. Nat. Commun. 11, 2155 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Jakob, R. P. et al. Dimeric structure of the bacterial extracellular foldase PrsA. J. Biol. Chem. 290, 3278–3292 (2015).

    Article  CAS  PubMed  Google Scholar 

  17. Stull, F., Betton, J.-M. & Bardwell, J. C. A. Periplasmic chaperones and prolyl isomerases. EcoSal Plus https://doi.org/10.1128/ecosalplus.ESP-0005-2018 (2018).

  18. Rouviere, P. E. & Gross, C. A. SurA, a periplasmic protein with peptidyl-prolyl isomerase activity, participates in the assembly of outer membrane porins. Genes Dev. 10, 3170–3182 (1996).

    Article  CAS  PubMed  Google Scholar 

  19. Xu, X., Wang, S., Hu, Y.-X. & McKay, D. B. The periplasmic bacterial molecular chaperone SurA adapts its structure to bind peptides in different conformations to assert a sequence preference for aromatic residues. J. Mol. Biol. 373, 367–381 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Hyyrylainen, H.-L. et al. Penicillin-binding protein folding is dependent on the PrsA peptidyl-prolyl cistrans isomerase in Bacillus subtilis. Mol. Microbiol. 77, 108–127 (2010).

    Article  CAS  PubMed  Google Scholar 

  21. Roch, M. et al. Thermosensitive PBP2a requires extracellular folding factors PrsA and HtrA1 for Staphylococcus aureus MRSA beta-lactam resistance. Commun. Biol. https://doi.org/10.1038/s42003-019-0667-0 (2019).

  22. Walton, T. A. & Sousa, M. C. Crystal structure of Skp, a prefoldin-like chaperone that protects soluble and membrane proteins from aggregation. Mol. Cell 15, 367–374 (2004).

    Article  CAS  PubMed  Google Scholar 

  23. Scholz, C. et al. SlyD proteins from different species exhibit high prolyl isomerase and chaperone activities. Biochemistry 45, 20–33 (2006).

    Article  CAS  PubMed  Google Scholar 

  24. Almeida, A. et al. A unified catalog of 204,938 reference genomes from the human gut microbiome. Nat. Biotechnol. 39, 105–114 (2021).

    Article  CAS  PubMed  Google Scholar 

  25. Li, J. et al. An integrated catalog of reference genes in the human gut microbiome. Nat. Biotechnol. 32, 834–841 (2014).

    Article  CAS  PubMed  Google Scholar 

  26. Elhenawy, W., Debelyy, M. O. & Feldman, M. F. Preferential packing of acidic glycosidases and proteases into Bacteroides outer membrane vesicles. mBio 5, e00909–e00914 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  27. Atanaskovic, I. & Kleanthous, C. Tools and approaches for dissecting protein bacteriocin import in gram-negative bacteria. Front. Microbiol. 10, 646 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Unal, C. M. & Steinert, M. Microbial peptidyl-prolyl cis/trans isomerases (PPIases): virulence factors and potential alternative drug targets. Microbiol. Mol. Biol. Rev. 78, 544–571 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  29. Scheuplein, N. J. et al. Targeting protein folding: a novel approach for the treatment of pathogenic bacteria. J. Med. Chem. 63, 13355–13388 (2020).

    Article  CAS  PubMed  Google Scholar 

  30. Bencivenga-Barry, N. A., Lim, B., Herrera, C. M., Trent, M. S. & Goodman, A. L. Genetic manipulation of wild human gut bacteroides. J. Bacteriol. https://doi.org/10.1128/jb.00544-19 (2020).

  31. García-Bayona, L. & Comstock, L. E. Streamlined genetic manipulation of diverse bacteroides and parabacteroides isolates from the human gut microbiota. mBio https://doi.org/10.1128/mBio.01762-19 (2019).

  32. Liu, D., Siguenza, N. E., Zarrinpar, A. & Ding, Y. Methods of DNA introduction for the engineering of commensal microbes. Eng. Microbiol. 2, 100048 (2022).

    Article  CAS  Google Scholar 

  33. Lim, B., Zimmermann, M., Barry, N. A. & Goodman, A. L. Engineered regulatory systems modulate gene expression of human commensals in the gut. Cell 169, 547–558.e515 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Jiang, K. et al. Functional characterization of Vip3Aa from Bacillus thuringiensis reveals the contributions of specific domains to its insecticidal activity. J. Biol. Chem. 299, 103000 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Jiang, K. et al. A strategy to enhance the insecticidal potency of Vip3Aa by introducing additional cleavage sites to increase its proteolytic activation efficiency. Eng. Microbiol. 3, 100083 (2023).

    Article  CAS  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Minor, W., Cymborowski, M., Otwinowski, Z. & Chruszcz, M. HKL-3000: the integration of data reduction and structure solution - from diffraction images to an initial model in minutes. Acta Crystallogr. D Struct. Biol. 62, 859–866 (2006).

    Article  Google Scholar 

  38. Liebschner, D. et al. Macromolecular structure determination using X-rays, neutrons and electrons: recent developments in Phenix. Acta Crystallogr. D Struct. Biol. 75, 861–877 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Emsley, P. & Cowtan, K. Coot: model-building tools for molecular graphics. Acta Crystallogr. D Struct. Biol. 60, 2126–2132 (2004).

    Article  Google Scholar 

  40. Camacho, C. et al. BLAST plus: architecture and applications. BMC Bioinform. https://doi.org/10.1186/1471-2105-10-421 (2009).

  41. 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 

  42. Minh, B. Q. et al. IQ-TREE 2: new models and efficient methods for phylogenetic inference in the genomic era. Mol. Biol. Evol. 37, 1530–1534 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Schmid, F. X. Prolyl isomerase: enzymatic catalysis of slow protein folding reactions. Annu. Rev. Biophys. Biomol. Struct. 22, 123–143 (1993).

    Article  CAS  PubMed  Google Scholar 

  44. Mucke, M. & Schmid, F. X. Intact disulfide bonds decelerate the folding of ribonuclease-T1. J. Mol. Biol. 239, 713–725 (1994).

    Article  CAS  PubMed  Google Scholar 

  45. Lian, H. et al. The Salmonella effector protein SopD targets Rab8 to positively and negatively modulate the inflammatory response. Nat. Microbiol 6, 658–671 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Spanò, S., Gao, X., Hannemann, S., Lara-Tejero, M. & Galán, J. E. A bacterial pathogen targets a host Rab-family GTPase defense pathway with a GAP. Cell Host Microbe 19, 216–226 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank A. L. Goodman (Yale University) for technical support and J. Galan (Yale University), Y. Shi (Westlake University) and H. Long (Ocean University of China) for discussion and careful review of the manuscript. We thank the staff from BL02U1 and BL18U1 beamlines of the National Facility for Protein Science Shanghai at SSRF for assistance during data collection. We also thank Y. Guo, J. Qu, J. Zhu, Z. Li, S. Wang, X. Zhao, H. Yu, X. Li and G. Lin from the core facilities for life and environmental sciences, SKLMT of Shandong University for their assistance in laser-scanning confocal microscopy, LC–MS/MS and X-ray diffraction experiments. The present study was supported by the National Natural Science Foundation of China (grant no. 32122007 to X.G.), the National Key R&D Program of China (grant no. 2022YFA1304200 to X.G.), the Shandong Provincial Natural Science Foundation (grant nos. ZR2021JQ09, ZR2019ZD21 and 2020CXGC011305 to X.G.) and the Youth Interdisciplinary Innovative Research Group of Shandong University (grant no. 2020QNQT009 to X.G.).

Author information

Authors and Affiliations

Authors

Contributions

X.G. and K.J. conceived the project. X.G., K.J., W.X.L., M.T., J.H.X., Z.C., C.L., B.L. and J.W.W. provided the methodology. X.G., K.J., W.X.L., M.T., J.H.X., Z.C., Y.Y, Y.R.Z. and X.Y.J. did the investigations. K.J., W.X.L., M.T., J.H.X. and Z.C. validated the results. X.G., K.J., W.X.L., M.T., Z.C. and F.S. wrote the original draft of the manuscript. X.G., K.J., W.X.L., D.L.W, B.L., F.S. and S.J.L. reviewed and edited the manuscript. X.G. acquired funding. X.G., X.Z.J, M.Y.W. and S.J.L. provided the resources. X.G. supervised the project.

Corresponding author

Correspondence to Xiang Gao.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Microbiology thanks Han Remaut, Kevin Foster 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 Screening of BfUbb sensitive B. fragilis strains.

a, Agar spot analysis is used to screen BfUbb-sensitive B. fragilis strains on a BfUbb-containing plate. 22 B. fragilis strains were spotted on a BHIS plate spread with 100 μg of BfUbb to evaluate their BfUbb sensitivity. The selected BfUbb-sensitive strains are listed on the right and underlined in red. b, Minimum inhibitory concentration (MIC) assays of BfUbb, tetracycline (tet), and chloramphenicol (cam) on B. fragilis GS077 and GS076. Strains were cultured in BHIS medium containing BfUbb, tet, or cam with the indicated gradient concentrations. Dotted lines represent the 50% concentrations of the control strains (CT). Data are expressed as the mean ± s.d. from three independent experiments. c, Western immunoblot analyses of BfUbb in the whole cell lysate (WCL) and culture supernatant (Sup) of strains of B. fragilis NCTC 9343 wildtype, ubb deleted mutant and ubb restored mutant. d, 5 µl culture supernatant of B. fragilis NCTC 9343 wildtype, ubb deleted mutant, and ubb restored mutant are spotted on the BHIS plates and then overlaid with a layer of agar containing the indicated strains. For a, c, and d, experiments were conducted at least three times with consistent results.

Source data

Extended Data Fig. 2 Morphology changes of B. fragilis GS077S or GS084I exposed to BfUbb.

a, b, Confocal microscopy images show the morphology of B. fragilis GS077S and GS084I exposed to BfUbb at concentrations of 0.1 μg/ml, 0.5 μg/ml, 1 μg/ml, and 10 μg/ml for 5 h. c, The linear correlation between BfUbb’s protein loading and band intensity was assessed through quantitative western blot analysis. The data are presented as the mean ± s.d. from at least three independent experiments, chosen to be representative of the overall trend. The true protein loading versus protein band intensity is depicted by the black curve, while the fitted linear relationship is represented by the dashed red line. The figure displays the specific mapping formula and coefficient of determination. d, Confocal microscopy images show the morphology of B. fragilis GS077S after being treated with the supernatant of B. fragilis NCTC 9343 for 4 h. e, Western immunoblot analysis of the amount of BfUbb secreted by ~4×107 B. fragilis NCTC 9343 (S1), the minimal amount of BfUbb needed to induce substantial elongation and growth defects in ~3.6×107B. fragilis GS077S (In), and the remaining BfUbb in the supernatant (S2). Data are representative of three independent experiments. f, Histograms showing the amount of BfUbb secreted by ~1×107 B. fragilis NCTC 9343 and the minimal amount of BfUbb needed to induce substantial elongation and growth defects in ~1×107B. fragilis GS077S. Data are expressed as the mean ± s.d. from three independent experiments. For b and d, Cellular membranes are stained with lipophilic dye FM 1-43 (green); scale bar, 5 μm; the images represent at least 10 images from three independent experiments.

Source data

Extended Data Fig. 3 Sequence alignment of PPIase proteins from B. fragilis NCTC 9343, GS077S, and GS084I.

Amino acid sequence alignment of the PPIase proteins from BfUbb-sensitive strain GS077S, BfUbb-insensitive strain GS084I, and BfUbb-encoding strain B. fragilis NCTC 9343.

Extended Data Fig. 4 PPIaseGS077 is the target of BfUbb.

a, b, Size-exclusion chromatography analyses of BfUbb in the presence of PPIaseGS077 (a) or PPIaseGS084 (b). Purified BfUbb was incubated with purified PPIaseGS077 or PPIaseGS084 and subsequently analysed using size-exclusion chromatography in a Superdex 75 increase column. Elution profiles along with SDS-PAGE analyses of the elution fractions are shown. This experiment was conducted at least three times with consistent results. c, d, f, The binding affinities of BfUbb (c and d) and BfUbbC70V (f) to PPIaseGS077 (c and f) or PPIaseGS084 (d) were measured by ITC. The N values were fixed to 1 during the data fitting for d and f. The Kd and ΔH values data are representative of 2-3 independent replicates. e, Confocal microscopy views of B. fragilis GS077S treated with BfUbb, BfUbbC70V, and BfUbbC76G after 5 hours, respectively. Cellular membranes are stained with lipophilic dye FM 1-43 (green); scale bar, 5 μm; the images represent at least 20 images from three independent experiments.

Source data

Extended Data Fig. 5 Tyr119 in PPIase is essential for interacting with BfUbb.

a, Confocal microscopy views of B. fragilis GS077S, GS077-PPIGS084(110-120), GS084I, and GS084-PPIGS077(110-120) treated with BfUbb after 5 hours. In strain GS077-PPIGS084(110-120), amino acids 110-120 of PPIaseGS077 were replaced with the corresponding amino acids 110-120 of PPIaseGS084. In strain GS084-PPIGS077(110-120), amino acids 110-120 of PPIaseGS084 were replaced with the corresponding amino acids 110-120 of PPIaseGS077. b, Confocal microscopy views of B. fragilis GS077S, GS077-PPIE118A/Y119E, GS084I, and GS084-PPIA118E/E119Y treated with BfUbb for 5 hours. c, Confocal microscopy views of B. fragilis GS077S, GS077-PPIE118A, GS077-PPIY119E, GS084I, GS084-PPIA118E, and GS084-PPIE119Y treated with BfUbb for 5 hours. For a, b, and c, cellular membranes are stained with lipophilic dye FM 1-43 (green); scale bars, 5 μm; the images represent at least 20 images from three independent experiments.

Extended Data Fig. 6 Crystal structures of the BfUbb–PPIaseGS077 complex and BfUbbC70V/C76G.

a, Structural alignment of the two copies of BfUbb–PPIaseGS077 complex in one asymmetric unit. The disulfide bond formed by Cys70 and Cys76 in BfUbb and the key interacting residue Y119 in PPIaseGS077 are shown as sticks. The black arrow indicates the angle of rotation around the central axis. b, c, The binding affinities of BfUbbD75G (b) and BfUbbN7A (c) to PPIaseGS077 were measured by ITC. The N values were fixed to 1 during the data fitting. The Kd and ΔH values data are representative of 2-3 independent replicates. d, Hydrophobic interactions between R8/I44/L68/I69/C70/C76 of BfUbb (light blue) and Y56/L55/L52 of PPIaseGS077 (green). e, f, B-factor diagram of BfUbb (e) and BfUbbC70V/C76G (f) represented by the B-factor putty program in PyMOL. The cartoon thickness and color reflect the relative Cα B-factors within the molecule. The disulfide bond between C70 and C76 in BfUbb is shown as a yellow stick.

Extended Data Fig. 7 The ratio of PPIase that needs to be targeted by BfUbb to see a cell growth defect.

a, The linear correlation between PPIase protein loading and band intensity was assessed through quantitative western blot analysis. The data are presented as the mean ± s.d. from at least three independent experiments. b, Confocal microscopy images of B. fragilis GS077S treated with BfUbb (0.15 μg/ml) after 3 h. c, Western immunoblot analysis of PPIase and BfUbb from ~8×107 cells of B. fragilis GS077S after incubation with BfUbb (0.15 μg/ml) for 3 h. Data are representative of at least three independent experiments. d, e, Histograms showing the amount (d) and concentration (e) of PPIase and BfUbb in ~1×107 cells of B. fragilis GS077S after they were treated with BfUbb (0.15 μg/ml) for 3 h. f, Confocal microscopy images of morphological changes in B. fragilis GS077ΔPPIaseGS077 harboring a plasmid expressing PPIaseGS077 in BHIS containing 0.1 mM rhamnose. The wild-type strain was used as the negative control. g, Western immunoblot analysis of the amount of PPIase in ~1×107 cells of B. fragilis GS077S and B. fragilis GS077ΔPPIaseGS077 harboring a plasmid expressing PPIaseGS077 in BHIS containing 0.1 mM rhamnose. Data are representative of three independent experiments. h, Histograms showing the relative concentration of PPIase in ~1×107 cells of B. fragilis GS077S and B. fragilis GS077ΔPPIaseGS077 harboring a plasmid expressing PPIaseGS077 in BHIS containing 0.1 mM rhamnose. The concentration of PPIase in B. fragilis GS077S is normalized as 100%. i, Co-immunoprecipitation analysis detecting the interaction of BfUbb with PPIase from B. fragilis GS077S when the majority of GS077S cells are exhibiting growth defects. j, Relative concentrations of BfUbb versus PPIase in Extended Data Fig. 7i. The concentration of PPIase is normalized as 100%. For b and f, Cellular membranes are stained with lipophilic dye FM 1-43 (green); scale bar, 5 μm; the images represent at least 10 images from three independent experiments. For d, e, h, and j, Data are expressed as the mean ± s.d. from three independent experiments.

Source data

Extended Data Fig. 8 Crystal structure of the PPIaseGS077.

a, Overall structure of PPIaseGS077. The two PPIaseGS077 molecules in one asymmetric unit are shown in cartoon models. b, c, Electron density maps showing the conformations of individual domains of the PPIaseGS077 in one asymmetric unit: chain A (b) and chain B (c). The 2Fo–Fc electron density maps of the chain A and chain B are contoured at 1.2 σ. d, e, PPIaseGS077 molecules in one asymmetric unit from the BfUbb–PPIaseGS077 complex are shown as cartoon models. The key residue Y119 is labeled and shown as a stick. f, g, PPIaseGS077 molecules in one asymmetric unit from the crystal structure PPIaseGS077 are shown as cartoon models. The key residue Y119 is highlighted in red. For a-g, NTD, N-terminal domain; CTD, C-terminal domain; P1 and P2, PPIase domain.

Extended Data Table 1 BfUbb-interacting proteins identified in GS077S

Supplementary information

Supplementary Information

Supplementary Fig. 1.

Reporting Summary

Supplementary Table 1

List of bacterial strains used in the present study.

Supplementary Table 2

Data collection and refinement statistics.

Supplementary Table 3

The LC–MS/MS analysis results of BfUbb-interacting proteins in GS077S.

Supplementary Table 4

Identification and co-occurrence of BfUbb, BfUbb-targeting PPIase, BfUbb-encoding strains, PPIase and BfUbb-insensitive PPIase homologues in human gut metagenomes EMBL.

Supplementary Table 5

Identification and co-occurrence of BfUbb, BfUbb-targeting PPIase, BfUbb-encoding strains, PPIase and BfUbb-insensitive PPIase homologues in human gut metagenomes 3CGC.

Supplementary Table 6

List of plasmids used in the present study.

Supplementary Table 7

Nucleotide sequences of PPIases used in thie present study.

Supplementary Video 1

Time-lapse of B. fragilis GS077S showing the morphology changes over the course of 8 s after being treated with BfUbb for 5 h. Cellular membranes are stained with lipophilic dye FM 1-43 (green). Scale bar, 10 μm. Taken under ×63 oil-immersion objective from three independent experiments.

Source data

Source Data Fig. 1

Uncropped plate corresponding to Fig. 1g.

Source Data Fig. 3

Uncropped plates, gels and western blots corresponding to Fig. 3c–e,h.

Source Data Fig. 3

Statistical source data for Fig. 3g.

Source Data Fig. 4

Uncropped plates and gels corresponding to Fig. 4b–g,l,m,o.

Source Data Fig. 5

Uncropped plates corresponding to Fig. 5a–c.

Source Data Fig. 5

Statistical source data for Fig. 5f–i.

Source Data Fig. 6

Statistical source data for Fig. 6a–f,h–l.

Source Data Extended Data Fig. 1

Uncropped western blots and gels corresponding to Extended Data Fig. 1c,d.

Source Data Extended Data Fig. 1

Statistical source data for Extended Data Fig. 1b.

Source Data Extended Data Fig. 2

Uncropped western blots corresponding to Extended Data Fig. 2c,e.

Source Data Extended Data Fig. 2

Statistical source data Extended Data for Fig. 2c,f.

Source Data Extended Data Fig. 4

Uncropped Coomassie Blue gels for Extended Data Fig. 4a,b.

Source Data Extended Data Fig. 4

Statistical source data for Extended Data Fig. 4a,b.

Source Data Extended Data Fig. 7

Uncropped western blots corresponding to Extended Data Fig. 7a,c,g,i.

Source Data Extended Data Fig. 7

Statistical source data for Extended Data Fig. 7a,d,e,h,j.

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

Jiang, K., Li, W., Tong, M. et al. Bacteroides fragilis ubiquitin homologue drives intraspecies bacterial competition in the gut microbiome. Nat Microbiol 9, 70–84 (2024). https://doi.org/10.1038/s41564-023-01541-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41564-023-01541-5

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