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:

A genetic switch controls Pseudomonas aeruginosa surface colonization

Abstract

Efficient colonization of mucosal surfaces is essential for opportunistic pathogens like Pseudomonas aeruginosa, but how bacteria collectively and individually adapt to optimize adherence, virulence and dispersal is largely unclear. Here we identified a stochastic genetic switch, hecR–hecE, which is expressed bimodally and generates functionally distinct bacterial subpopulations to balance P. aeruginosa growth and dispersal on surfaces. HecE inhibits the phosphodiesterase BifA and stimulates the diguanylate cyclase WspR to increase c-di-GMP second messenger levels and promote surface colonization in a subpopulation of cells; low-level HecE-expressing cells disperse. The fraction of HecE+ cells is tuned by different stress factors and determines the balance between biofilm formation and long-range cell dispersal of surface-grown communities. We also demonstrate that the HecE pathway represents a druggable target to effectively counter P. aeruginosa surface colonization. Exposing such binary states opens up new ways to control mucosal infections by a major human pathogen.

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: The hecRE module regulates biofilm formation by altering c-di-GMP levels.
Fig. 2: HecE inhibits the phosphodiesterase BifA and activates the diguanylate cyclase WspR.
Fig. 3: The expression of hecR and hecE is autoregulated, bimodal and tuned by stress.
Fig. 4: HecE controls P. aeruginosa biofilm formation and dispersal.
Fig. 5: HecE and H6-335-P1 antagonistically control BifA activity.

Similar content being viewed by others

Data availability

The datasets generated and/or analysed during this study are available from the corresponding author on reasonable request. The raw sequencing files of the chromatin immunoprecipitation with sequencing experiment can be accessed at the NCBI under the accession number PRJNA900431. Unique biological materials are available from the corresponding author on reasonable request. The structural coordinates of the BifA R-state dimer are deposited in the PDB library under the accession number 8ARV.

Code availability

The code generated for the analysis of flow cytometry data can be accessed at https://github.com/Jenal-Lab.

References

  1. Rutherford, S. T. & Bassler, B. L. Bacterial quorum sensing: its role in virulence and possibilities for its control. CSH Perspect. Med. 2, a012427 (2012).

    Google Scholar 

  2. Rahbari, K. M., Chang, J. C. & Federle, M. J. A Streptococcus quorum sensing system enables suppression of innate immunity. mBio 12, e03400-20 (2021).

    PubMed  PubMed Central  Google Scholar 

  3. Wu, L. & Luo, Y. Bacterial quorum-sensing systems and their role in intestinal bacteria-host crosstalk. Front. Microbiol. 12, 611413 (2021).

    PubMed  PubMed Central  Google Scholar 

  4. Laventie, B.-J. et al. A surface-induced asymmetric program promotes tissue colonization by Pseudomonas aeruginosa. Cell Host Microbe 25, 140–152 (2019).

    CAS  PubMed  Google Scholar 

  5. Diard, M. et al. Stabilization of cooperative virulence by the expression of an avirulent phenotype. Nature 494, 353–356 (2013).

    CAS  PubMed  Google Scholar 

  6. Ackermann, M. et al. Self-destructive cooperation mediated by phenotypic noise. Nature 454, 987–990 (2008).

    CAS  PubMed  Google Scholar 

  7. Kotte, O., Volkmer, B., Radzikowski, J. L. & Heinemann, M. Phenotypic bistability in Escherichia coli’s central carbon metabolism. Mol. Syst. Biol. 10, 736 (2014).

    PubMed  PubMed Central  Google Scholar 

  8. Basan, M. et al. A universal trade-off between growth and lag in fluctuating environments. Nature 584, 470–474 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Bakshi, S. et al. Tracking bacterial lineages in complex and dynamic environments with applications for growth control and persistence. Nat. Microbiol. 6, 783–791 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Balaban, N. Q., Merrin, J., Chait, R., Kowalik, L. & Leibler, S. Bacterial persistence as a phenotypic switch. Science 305, 1622–1625 (2004).

    CAS  PubMed  Google Scholar 

  11. Arnoldini, M., Vizcarra, I. A. & Peña-Miller, R. Bistable expression of virulence genes in Salmonella leads to the formation of an antibiotic-tolerant subpopulation. PLoS Biol. 8, e1001928 (2014).

    Google Scholar 

  12. Manina, G., Griego, A., Singh, L. K., McKinney, J. D. & Dhar, N. Preexisting variation in DNA damage response predicts the fate of single mycobacteria under stress. EMBO J. 38, e101876 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Stewart, M. K., Cummings, L. A., Johnson, M. L., Berezow, A. B. & Cookson, B. T. Regulation of phenotypic heterogeneity permits Salmonella evasion of the host caspase-1 inflammatory response. Proc. Natl Acad. Sci. USA 108, 20742–20747 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Kim, J. M., Garcia-Alcala, M., Balleza, E. & Cluzel, P. Stochastic transcriptional pulses orchestrate flagellar biosynthesis in Escherichia coli. Sci. Adv. 6, eaax0947 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Armbruster, C. R. et al. Heterogeneity in surface sensing suggests a division of labor in Pseudomonas aeruginosa populations. eLife 8, e45084 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Klockgether, J. & Tümmler, B. Recent advances in understanding Pseudomonas aeruginosa as a pathogen. F1000Research 6, 1261 (2017).

    PubMed  PubMed Central  Google Scholar 

  17. Kazmierczak, B. I., Schniederberend, M. & Jain, R. Cross-regulation of Pseudomonas motility systems: the intimate relationship between flagella, pili and virulence. Curr. Opin. Microbiol. 28, 78–82 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Malone, J. G. et al. YfiBNR mediates cyclic di-GMP dependent small colony variant formation and persistence in Pseudomonas aeruginosa. PLoS Pathog. 6, e1000804 (2010).

    PubMed  PubMed Central  Google Scholar 

  19. Wang, C., Ye, F., Kumar, V., Gao, Y.-G. & Zhang, L.-H. BswR controls bacterial motility and biofilm formation in Pseudomonas aeruginosa through modulation of the small RNA rsmZ. Nucleic Acids Res. 42, 4563–4576 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Wurtzel, O. et al. The single-nucleotide resolution transcriptome of Pseudomonas aeruginosa grown in body temperature. PLoS Pathog. 8, e1002945 (2012).

    PubMed  PubMed Central  Google Scholar 

  21. Sevin, E. W. & Barloy-Hubler, F. RASTA-Bacteria: a web-based tool for identifying toxin–antitoxin loci in prokaryotes. Genome Biol. 8, R155 (2007).

    PubMed  PubMed Central  Google Scholar 

  22. Kaczmarczyk, A. et al. A novel biosensor reveals dynamic changes of C-di-GMP in differentiating cells with ultra-high temporal resolution. Preprint at bioRxiv https://doi.org/10.1101/2022.10.18.512705 (2022).

  23. Harrison, J. J. et al. Elevated exopolysaccharide levels in Pseudomonas aeruginosa flagellar mutants have implications for biofilm growth and chronic infections. PLoS Genet. 16, e1008848 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Dreifus, J. E. et al. The Sia System and c-di-GMP play a crucial role in controlling cell-association of Psl in planktonic P. aeruginosa. J. Bacteriol. 204, e0033522 (2022).

    PubMed  Google Scholar 

  25. Kuchma, S. L. et al. BifA, a cyclic-di-GMP phosphodiesterase, inversely regulates biofilm formation and swarming motility by Pseudomonas aeruginosa PA14. J. Bacteriol. 189, 8165–8178 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Hickman, J. W., Tifrea, D. F. & Harwood, C. S. A chemosensory system that regulates biofilm formation through modulation of cyclic diguanylate levels. Proc. Natl Acad. Sci. USA 102, 14422–14427 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Aldridge, P., Paul, R., Goymer, P., Rainey, P. & Jenal, U. Role of the GGDEF regulator PleD in polar development of Caulobacter crescentus. Mol. Microbiol. 47, 1695–1708 (2003).

    CAS  PubMed  Google Scholar 

  28. Andersen, J. B. et al. Induction of native c-di-GMP phosphodiesterases leads to dispersal of Pseudomonas aeruginosa biofilms. Antimicrob. Agents Chem. 65, e02431-20 (2021).

    Google Scholar 

  29. McAdams, H. H. & Arkin, A. Stochastic mechanisms in gene expression. Proc. Natl Acad. Sci. USA 94, 814–819 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Brencic, A. et al. The GacS/GacA signal transduction system of Pseudomonas aeruginosa acts exclusively through its control over the transcription of the RsmY and RsmZ regulatory small RNAs. Mol. Microbiol. 73, 434–445 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Francis, V. I., Stevenson, E. C. & Porter, S. L. Two-component systems required for virulence in Pseudomonas aeruginosa. FEMS Microbiol. Lett. 364, fnx104 (2017).

    PubMed  PubMed Central  Google Scholar 

  32. Wang, B. X. et al. Mucin glycans signal through the sensor kinase rets to inhibit virulence-associated traits in Pseudomonas aeruginosa. Curr. Biol. 31, 90–102 (2021).

    CAS  PubMed  Google Scholar 

  33. Broder, U. N., Jaeger, T. & Jenal, U. LadS is a calcium-responsive kinase that induces acute-to-chronic virulence switch in Pseudomonas aeruginosa. Nat. Microbiol. 2, 16184 (2016).

    CAS  PubMed  Google Scholar 

  34. LeRoux, M., Peterson, S. B. & Mougous, J. D. Bacterial danger sensing. J. Mol. Biol. 427, 3744–3753 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Palmer, K. L., Aye, L. M. & Whiteley, M. Nutritional cues control Pseudomonas aeruginosa multicellular behavior in cystic fibrosis sputum. J. Bacteriol. 189, 8079–8087 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Sriramulu, D. D., Lünsdorf, H., Lam, J. S. & Römling, U. Microcolony formation: a novel biofilm model of Pseudomonas aeruginosa for the cystic fibrosis lung. J. Med. Microbiol. 54, 667–676 (2005).

    PubMed  Google Scholar 

  37. Irie, Y. et al. Self-produced exopolysaccharide is a signal that stimulates biofilm formation in Pseudomonas aeruginosa. Proc. Natl Acad. Sci. USA 109, 20632–20636 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Andersen, J. B. et al. Identification of small molecules that interfere with c-di-GMP signaling and induce dispersal of Pseudomonas aeruginosa biofilms. npj Biofilms Microbiomes 7, 59 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Sundriyal, A. et al. Inherent regulation of EAL domain-catalyzed hydrolysis of second messenger cyclic di-GMP. J. Biol. Chem. 289, 6978–6990 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Reinders, A. et al. Digital control of c-di-GMP in E. coli balances population-wide developmental transitions and phage sensitivity. Preprint at bioRxiv https://doi.org/10.1101/2021.10.01.462762 (2021).

  41. Kulesekara, H., Lee, V. & Brencic, A. Analysis of Pseudomonas aeruginosa diguanylate cyclases and phosphodiesterases reveals a role for bis-(3′-5′)-cyclic-GMP in virulence. Proc. Natl Aacd. Sci. USA 103, 2839–2844 (2006).

    CAS  Google Scholar 

  42. Melaugh, G. et al. Distinct types of multicellular aggregates in Pseudomonas aeruginosa liquid cultures. Preprint at bioRxiv https://doi.org/10.1101/2022.05.18.492589 (2022).

  43. Roy, A. B., Petrova, O. E. & Sauer, K. The phosphodiesterase DipA (PA5017) is essential for Pseudomonas aeruginosa biofilm dispersion. J. Bacteriol. 194, 2904–2915 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Silva, J. B., Storms, Z. & Sauvageau, D. Host receptors for bacteriophage adsorption. FEMS Microbiol. Lett. 363, fnw002 (2016).

    Google Scholar 

  45. Harvey, H. et al. Pseudomonas aeruginosa defends against phages through type IV pilus glycosylation. Nat. Microbiol. 3, 47–52 (2018).

    CAS  PubMed  Google Scholar 

  46. Ackermann, M. A functional perspective on phenotypic heterogeneity in microorganisms. Nat. Rev. Microbiol. 13, 497–508 (2015).

    CAS  PubMed  Google Scholar 

  47. Lowery, N. V., McNally, L., Ratcliff, W. C. & Brown, S. P. Division of labor, bet hedging, and the evolution of mixed biofilm investment strategies. mBio 8, e00672-17 (2017).

    PubMed  PubMed Central  Google Scholar 

  48. Almblad, H. et al. The cyclic AMP–Vfr signaling pathway in Pseudomonas aeruginosa is inhibited by cyclic di-GMP. J. Bacteriol. 197, 2190–2200 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Kulasekara, B. R. et al. c-di-GMP heterogeneity is generated by the chemotaxis machinery to regulate flagellar motility. eLife 2, e01402 (2013).

    PubMed  PubMed Central  Google Scholar 

  50. Burrows, L. L. Pseudomonas aeruginosa twitching motility: type IV pili in action. Annu. Rev. Microbiol. 66, 493–520 (2012).

    CAS  PubMed  Google Scholar 

  51. Ha, D.-G. & O’Toole, G. A. c-di-GMP and its effects on biofilm formation and dispersion: a Pseudomonas aeruginosa review. Microbiol. Spectr. 3, MB-0003-2014 (2015).

    PubMed  Google Scholar 

  52. Vrla, G. D. et al. Cytotoxic alkyl-quinolones mediate surface-induced virulence in Pseudomonas aeruginosa. PLoS Pathog. 16, e1008867 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Mulcahy, L. R., Isabella, V. M. & Lewis, K. Pseudomonas aeruginosa biofilms in disease. Microb. Ecol. 68, 1–12 (2014).

    CAS  PubMed  Google Scholar 

  54. Siryaporn, A., Kuchma, S. L., O’Toole, G. A. & Gitai, Z. Surface attachment induces Pseudomonas aeruginosa virulence. Proc. Natl Acad. Sci. USA 111, 16860–16865 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Acar, M., Mettetal, J. T. & van Oudenaarden, A. Stochastic switching as a survival strategy in fluctuating environments. Nat. Genet. 40, 471–475 (2008).

    CAS  PubMed  Google Scholar 

  56. Perkins, T. J. & Swain, P. S. Strategies for cellular decision‐making. Mol. Syst. Biol. 5, 326–326 (2009).

    PubMed  PubMed Central  Google Scholar 

  57. Neidhardt, F. C., Bloch, P. L. & Smith, D. F. Culture medium for enterobacteria. J. Bacteriol. 119, 736–747 (1974).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Maffei, E. et al. Systematic exploration of Escherichia coli phage–host interactions with the BASEL phage collection. PLoS Biol. 19, e3001424 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Smyth, G. K. Linear models and empirical bayes methods for assessing differential expression in microarray experiments. Stat. Appl Genet Mol. 3, Article 3 (2004).

    Google Scholar 

  60. Agustoni, E. et al. Acquisition of enzymatic progress curves in real time by quenching-free ion exchange chromatography. Anal. Biochem. 639, 114523 (2022).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Potterton, L. et al. CCP4i2: the new graphical user interface to the CCP4 program suite. Acta Crystallogr. D 74, 68–84 (2018).

    CAS  Google Scholar 

  63. McCoy, A. J. et al. Phaser crystallographic software. J. Appl. Crystallogr. 40, 658–674 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Eswar, N. et al. Comparative protein structure modeling using modeller. Curr. Protoc. Bioinform. 15, 5.6.1–5.6.30 (2006).

    Google Scholar 

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

    PubMed  Google Scholar 

  66. Afonine, P. V. et al. Real-space refinement in PHENIX for cryo-EM and crystallography. Acta Crystallogr. D 74, 531–544 (2018).

    CAS  Google Scholar 

  67. Pettersen, E. F. et al. UCSF ChimeraX: structure visualization for researchers, educators, and developers. Protein Sci. 30, 70–82 (2020).

    PubMed  PubMed Central  Google Scholar 

  68. Wilson, D. S. & Keefe, A. D. Random mutagenesis by PCR. Curr. Protoc. Mol. Biol. 51, 8.3.1–8.3.9 (2000).

    Google Scholar 

  69. Newman, J. R. & Fuqua, C. Broad-host-range expression vectors that carry the l-arabinose-inducible Escherichia coli araBAD promoter and the araC regulator. Gene 227, 197–203 (1999).

    CAS  PubMed  Google Scholar 

  70. Hartmann, R. et al. Quantitative image analysis of microbial communities with BiofilmQ. Nat. Microbiol. 6, 151–156 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Budzik, J. M., Rosche, W. A., Rietsch, A. & O’Toole, G. A. Isolation and characterization of a generalized transducing phage for Pseudomonas aeruginosa strains PAO1 and PA14. J. Bacteriol. 186, 3270–3273 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank F. Hamburger (Biozentrum, University of Basel) for her help with cloning. We thank E. Maffei and A. Harms (Biozentrum, University of Basel) for their support with phage isolation and characterization. This work was supported by a Biozentrum PhD Fellowship to C.M., the Swiss National Science Foundation (grant no. 310030_189253 to U.J.), the NCCR AntiResist funded by Swiss National Science Foundation (grant no. 51NF40_180541 to K.D. and U.J.), the European Research Council (grant no. 716734 to K.D.) and grants from Sygeforsikring Danmark, Danish Innovation Fund and Novoseed to M.G. and T.T.-N.

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization: C.M., T.J. and U.J. Methodology: C.M., R.D.T., D.S., A.K., R.Z., T.J., B.-J.L., J.G.M., J.B.A., M.D., T.T.-N., K.Q., K.D. and U.J. Formal analysis: C.M., R.D.T., S.B. and D.S. Investigation: C.M., R.D.T., D.S., R.Z., T.J., J.G.M., F.W. and J.B.A. Writing (original draft): C.M. and U.J. Funding acquisition: M.G., T.T.-N., K.D., K.Q., S.H. and U.J. Supervision: S.H., K.D. and U.J.

Corresponding author

Correspondence to Urs Jenal.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Microbiology thanks Tim Rick, Daniel Wozniak 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 The hecRE module regulates SCV and biofilm formation by altering c-di-GMP levels.

a, Colony morphology of the hecR::Tn mutant. b, Colony morphology of WT and hec deletion strains. Experiments were performed in triplicate; one representative image is shown. c, Growth of P. aeruginosa WT and hec deletion strains with mean values and the s.d. of three biological replicates (with six technical replicates each). d, Growth of WT and hec deletion strains containing plasmids for expression of hec genes from an IPTG-inducible promoter (EV, control plasmid). Growth in LB (dashed lines) or LB supplemented with 100 µM IPTG (solid lines) was recorded, and mean values and the s.d. of three biological replicates (with three technical replicates each) are shown. e, Colony morphology of P. aeruginosa WT and mutants lacking Pel or Psl exopolysaccharides expressing hecE from an IPTG-inducible promoter. Experiments were done in triplicate; one representative image is shown.

Extended Data Fig. 2 HecE modulates c-di-GMP levels by inhibiting the phosphodiesterase BifA and activating the diguanylate cyclase WspR.

a, Schematic of the BifA domain architecture and BifA fragment identified by Y2H to specifically interact with HecE (TM, transmembrane domain; SID, smallest interacting domain). b, Three-dimensional structure of a BifA dimer embedded in the cytoplasmic membrane (grey) as predicted by AlphaFold. c, CoIP-MS analysis of HecE. Immunoprecipitation experiments were performed using the PAO1 WT strain or a strain expressing C-terminally FLAG-tagged HecE and anti-M2-conjugated magnetic beads. Proteins retained on the beads were analysed using mass spectrometry. Data obtained from three biological replicates are shown as volcano plots. Log2-transformed intensity ratios of the detected peptides between HecE–Flag and the WT (ctrl) were calculated and plotted versus values derived from significance analysis (modified t-statistic, empirical Bayes method59). d, Colony morphology of P. aeruginosa wild-type and mutant strains containing an empty plasmid (EV) or a plasmid expressing hecE from an IPTG-inducible promoter. Experiments were done in triplicate; one representative image is shown. e, Growth (top) and fluorescence (c-di-GMP levels, bottom) of P. aeruginosa wild-type (WT), hecR::Tn and ΔbifA strains carrying a plasmid control (EV) or a plasmid expressing the c-di-GMP sensor cdGreen. Mean values and the s.d. of two biological replicates (with six technical replicates each) are shown.

Extended Data Fig. 3 Expression of hecR and hecE is autoregulated, bimodal and tuned by stress.

a, Expression of hecR boosts HecR and HecE levels. Immunoblot analysis of HecR and HecE in strains expressing chromosomal Flag-tagged copies of hecR or hecE. Strains contained a plasmid expressing hecR, hecE or both genes from an arabinose-inducible promoter (EV, control plasmid). Extracts of cells grown in the presence (+) or absence (−) of the inducer arabinose were analysed by immunobloting with an anti-Flag (n = 1). b, HecR binds to the hecRE promoter region. EMSA assays were carried out with labelled DNA containing the hecRE promoter region and with increasing concentrations of purified HecR protein, as indicated (n = 2). c, HecR exclusively binds to the hecRE locus on the P. aeruginosa chromosome. Chromatin pull-down experiments were performed with a strain expressing a HA-tagged copy of hecR in the absence of the chromosomal wild-type hecR copy and with HA-specific antibodies. Sequencing reads are plotted on the y axis for the entire P. aeruginosa chromosome (top) or the hecRE locus (bottom). Graphs were generated using Geneious version 2019.0 created by Biomatters. d, The fraction of cells expressing hecE changes during P. aeruginosa growth. The optical density (solid lines) and fluorescence (hecE expression; dashed lines) were recorded during the growth of hecRE-2mR reporter strains containing a plasmid expressing hecR, hecE or both genes from an IPTG-inducible promoter (EV, control plasmid). Mean values and the s.d. are shown for three biological replicates with three technical replicates each. e, The fraction of cells expressing hecRE is independent of the culture medium. P. aeruginosa wild-type carrying the hecRE-2mR chromosomal reporter was cultured in the indicated medium for 20 h. Quantification of flow cytometry data is shown for three biological replicates. Triplicates were fitted independently. Mean and range of the calculated values are shown. Normality of the means was checked usinf a Shapiro–Wilk test (P < 0.05). A one-way ANOVA test was used to analyse differences in population means (with P < 0.05), followed by Tukey’s HSD post-hoc test, for which adjusted P values are shown. f, The fraction of cells expressing hecRE varies during P. aeruginosa growth. hecRE-2mR reporter strains carrying a plasmid expressing hecR from an IPTG-inducible promoter were diluted back into fresh LB and cultured for 12 h (EV, control plasmid). Fluorescence intensities of the reporter, as measured by flow cytometry, are shown as violin plots (left y axis) and growth is indicated in stippled lines (right y axis). Cells with signal values higher than the indicated threshold (grey line) were counted as cells expressing hecRE-2mR. g, Expression of hecRE responds to nutrient limitations. The P. aeruginosa hecRE-2mR reporter strain was cultured in MOPS minimal medium with different amounts of succinate as the sole carbon source. Growth and the fraction of cells expressing hecRE are indicated as in Fig. 3f. h,i, Flow cytometry data of the hecRE-2mR reporter in ΔrsmA (h) and ΔrpoS (i) mutants, cultured in MOPS with the indicated amounts of succinate at 37 °C. Cells with higher fluorescence intensity than the PAO1 wild-type strain were quantified (right axis). The optical density of the culture before flow cytometry is indicated in stippled lines (left axis). j, Ectopic expression of hecR induces hecE transcription only in the stationary phase. Flow cytometry analysis of strains with the hecRE-2mR reporter in the wild type (WT) and indicated mutants. Reporter strains containing a plasmid expressing hecR were cultured in LB supplemented with 100 µM IPTG either to exponential phase or for 20 h (stationary). Flow cytometry experiments are shown for three biological replicates with the individual histograms stacked in the graph. k, The fraction of cells expressing hecRE is increased at elevated temperature. The P. aeruginosa hecRE-2mR reporter strain harbouring a plasmid expressing hecR from an IPTG-inducible promoter was culktured in MOPS + 20 mM succinate supplemented with 100 µM IPTG at increasing temperatures, as indicated (EV, control plasmid). Flow cytometry experiments are shown for three biological replicates with the individual histograms stacked in the graph. l, Uncropped scan of the immunoblot shown in a. m, Gating strategy used for the flow cytometry experiments. Shown is one of three replicates of the wild type carrying the hecRE-2mR reporter.

Extended Data Fig. 4 Heterogeneous expression of hecE controls surface motility, biofilm formation and dispersal.

a,d, Increase of biofilm volume of P. aeruginosa wild-type and mutant strains. Biofilms were segmented into cubes with a side length of 2.34 µm using BiofilmQ70 and biofilm volumes were calculated as the sum of all individual cubes at any given time. The red line indicates the start of the dispersal event. Analysis for one experiment is shown in each panel. b,e, Object parameters including the distance to biofilm boundary (resolution, 20 voxels), fluorescent intensities and local cell density were calculated using BiofilmQ. c,f, Fluorescent intensities at the start of the dispersal event (red line) were plotted and colour-coded according to their distance to the biofilm boundary.

Extended Data Fig. 5 HecE and H6-335-P1 antagonistically control BifA activity by influencing the T-to-R-state equilibrium.

a, Standard deviation of the mean of the attachment data shown in Fig. 5a. b, Microtiter plate-based attachment of WT and mutant strains carrying a plasmid expressing the E. coli diguanylate cyclase dgcZ from an IPTG-inducible promoter and various bifA alleles from a cumate-inducible promoter (EV, control plasmid). BifA expression was not induced, dgcZ expression was induced with 100 µM IPTG. H6-335-P1 was added at the indicated concentrations. The attachment was quantified after 9.5 h. Mean values of two biological replicates are shown. c, Arrangement of α6 and α5′ of the PdeL dimer in the R- (PDB: 4LYK) and T-state (PDB: 4LJ3). Residues of the active site and residues involved in the stabilization of the T-state conformation are highlighted and shown as a stick representation. d, Arrangement of α6 and α5′ of the BifA dimer in the R- (PDB: 8ARV) and modelled T-state. Residues of the active site and conserved residues postulated to stabilize the T-state conformation are highlighted and shown as a stick representation. e,f, Standard deviation of the mean of the attachment data shown in Fig. 5d,e, respectively.

Supplementary information

Reporting Summary

Supplementary Table 1

SCV screen

Supplementary Table 2

Yeast-two-hybrid screen

Supplementary Table 3

SCV suppressor screen

Supplementary Table 4

Structure data collection and refinement statistics

Supplementary Video 1

P. aeruginosa cells expressing hecE remain attached after biofilm dispersal.

Supplementary Video 2

Cell dispersal from flow chamber-grown biofilms is a highly coordinated process.

Supplementary Video 3

Mutants lacking HecE fail to maintain a mature biofilm after cell dispersal.

Supplementary Video 4

Mutants lacking BifA undergo premature biofilm formation and cell dispersal.

Supplementary Video 5

Mutants lacking WspR fail to maintain a mature biofilm after cell dispersal.

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

Manner, C., Dias Teixeira, R., Saha, D. et al. A genetic switch controls Pseudomonas aeruginosa surface colonization. Nat Microbiol 8, 1520–1533 (2023). https://doi.org/10.1038/s41564-023-01403-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41564-023-01403-0

This article is cited by

Search

Quick links

Nature Briefing Microbiology

Sign up for the Nature Briefing: Microbiology newsletter — what matters in microbiology research, free to your inbox weekly.

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