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:

Porin-independent accumulation in Pseudomonas enables antibiotic discovery

Abstract

Gram-negative antibiotic development has been hindered by a poor understanding of the types of compounds that can accumulate within these bacteria1,2. The presence of efflux pumps and substrate-specific outer-membrane porins in Pseudomonas aeruginosa renders this pathogen particularly challenging3. As a result, there are few antibiotic options for P. aeruginosa infections4 and its many porins have made the prospect of discovering general accumulation guidelines seem unlikely5. Here we assess the whole-cell accumulation of 345 diverse compounds in P. aeruginosa and Escherichia coli. Although certain positively charged compounds permeate both bacterial species, P. aeruginosa is more restrictive compared to E. coli. Computational analysis identified distinct physicochemical properties of small molecules that specifically correlate with P. aeruginosa accumulation, such as formal charge, positive polar surface area and hydrogen bond donor surface area. Mode of uptake studies revealed that most small molecules permeate P. aeruginosa using a porin-independent pathway, thus enabling discovery of general P. aeruginosa accumulation trends with important implications for future antibiotic development. Retrospective antibiotic examples confirmed these trends and these discoveries were then applied to expand the spectrum of activity of a gram-positive-only antibiotic, fusidic acid, into a version that demonstrates a dramatic improvement in antibacterial activity against P. aeruginosa. We anticipate that these discoveries will facilitate the design and development of high-permeating antipseudomonals.

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: Assessment of antibiotic controls in P. aeruginosa accumulation assay and evaluation of eNTRy rules predictability in P. aeruginosa.
Fig. 2: Importance of ClogD7.4, HBD surface area and positive charge descriptors for accumulation in P. aeruginosa PAO1.
Fig. 3: Evaluation of accumulation in a porin-deficient strain and in the presence of magnesium.
Fig. 4: Six retrospective examples in which an antibiotic derivative is more active (more than fourfold) against P. aeruginosa than the parent.
Fig. 5: Development of PA-active FA derivative.

Similar content being viewed by others

Data availability

All data supporting the findings of this study are available in the paper and the Supplementary Information.  Source data are provided with this paper.

Code availability

Source code for data analysis and model training can be found at https://github.com/HergenrotherLab/GramNegAccum.

References

  1. Silver, L. L. A Gestalt approach to Gram-negative entry. Bioorg. Med. Chem. 24, 6379–6389 (2016).

    Article  CAS  PubMed  Google Scholar 

  2. Payne, D. J., Gwynn, M. N., Holmes, D. J. & Pompliano, D. L. Drugs for bad bugs: confronting the challenges of antibacterial discovery. Nat. Rev. Drug. Discov. 6, 29–40 (2007).

    Article  CAS  PubMed  Google Scholar 

  3. Yoshimura, F. & Nikaido, H. Permeability of Pseudomonas aeruginosa outer membrane to hydrophilic solutes. J. Bacteriol. 152, 636–642 (1982).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Bassetti, M., Vena, A., Croxatto, A., Righi, E. & Guery, B. How to manage Pseudomonas aeruginosa infections. Drugs Context 7, 212527 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  5. Tommasi, R., Iyer, R. & Miller, A. A. Antibacterial drug discovery: some assembly required. ACS Infect. Dis. 4, 686–695 (2018).

    Article  CAS  PubMed  Google Scholar 

  6. Tamber, S., Ochs, M. M. & Hancock, R. E. Role of the novel OprD family of porins in nutrient uptake in Pseudomonas aeruginosa. J. Bacteriol. 188, 45–54 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Aeschlimann, J. R. The role of multidrug efflux pumps in the antibiotic resistance of Pseudomonas aeruginosa and other Gram-negative bacteria. Insights from the Society of Infectious Diseases Pharmacists. Pharmacotherapy 23, 916–924 (2003).

    Article  CAS  PubMed  Google Scholar 

  8. Krishnamoorthy, G. et al. Synergy between active efflux and outer membrane diffusion defines rules of antibiotic permeation into Gram-negative bacteria. mBio 8, e01172–17 (2017).

  9. Falagas, M. E. & Kasiakou, S. K. Toxicity of polymyxins: a systematic review of the evidence from old and recent studies. Crit. Care 10, R27 (2006).

    Article  PubMed  PubMed Central  Google Scholar 

  10. Mingeot-Leclercq, M. P. & Tulkens, P. M. Aminoglycosides: nephrotoxicity. Antimicrob. Agents Chemother. 43, 1003–1012 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Richter, M. F. & Hergenrother, P. J. The challenge of converting Gram-positive-only compounds into broad-spectrum antibiotics. Ann. N. Y. Acad. Sci. 1435, 18–38 (2019).

    Article  ADS  CAS  PubMed  Google Scholar 

  12. Zgurskaya, H. I. & Rybenkov, V. V. Permeability barriers of Gram-negative pathogens. Ann. N. Y. Acad. Sci. 1459, 5–18 (2020).

    Article  ADS  PubMed  Google Scholar 

  13. Richter, M. F. et al. Predictive compound accumulation rules yield a broad-spectrum antibiotic. Nature 545, 299–304 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  14. Motika, S. E. et al. Antibiotic active through inhibition of an essential riboswitch. J. Am. Chem. Soc. 142, 10856–10862 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Parker, E. N. et al. Implementation of permeation rules leads to a FabI inhibitor with activity against Gram-negative pathogens. Nat. Microbiol. 5, 67–75 (2020).

    Article  CAS  PubMed  Google Scholar 

  16. Hu, Y. et al. Discovery of pyrido[2,3-b]indole derivatives with Gram-negative activity targeting both DNA gyrase and topoisomerase IV. J. Med. Chem. 63, 9623–9649 (2020).

    Article  CAS  PubMed  Google Scholar 

  17. Andrews, L. D. et al. Optimization and mechanistic characterization of pyridopyrimidine inhibitors of bacterial biotin carboxylase. J. Med. Chem. 62, 7489–7505 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Lukežič, T. et al. Engineering atypical tetracycline formation in Amycolatopsis sulphurea for the production of modified chelocardin antibiotics. ACS Chem. Biol. 14, 468–477 (2019).

    Article  PubMed  Google Scholar 

  19. Skepper, C. K. et al. Topoisomerase inhibitors addressing fluoroquinolone resistance in Gram-negative bacteria. J. Med. Chem. 63, 7773–7816 (2020).

    Article  CAS  PubMed  Google Scholar 

  20. Brem, J. et al. Imitation of beta-lactam binding enables broad-spectrum metallo-beta-lactamase inhibitors. Nat. Chem. 14, 15–24 (2022).

    Article  CAS  PubMed  Google Scholar 

  21. Schumacher, C. E. et al. Total synthesis and antibiotic properties of amino-functionalized aromatic terpenoids related to erogorgiaene and the pseudopterosins. Eur. J. Org. Chem. 2022, e202200058 (2022).

  22. Parker, E. N. et al. An iterative approach guides discovery of the FabI inhibitor fabimycin, a late-stage antibiotic candidate with in vivo efficacy against drug-resistant Gram-negative infections. ACS Cent. Sci. 8, 1145–1158 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Huang, K.-J. et al. Deletion of a previously uncharacterized lipoprotein lirL confers resistance to an inhibitor of type II signal peptidase in Acinetobacter baumannii. Proc. Natl Acad. Sci. USA 119, e2123117119 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Onyedibe, K. I. et al. Re-sensitization of multidrug-resistant and colistin-resistant Gram-negative bacteria to colistin by Povarov/Doebner-derived compounds. ACS Infect. Dis. 9, 283–295 (2023).

    Article  CAS  PubMed  Google Scholar 

  25. Goethe, O., DiBello, M. & Herzon, S. B. Total synthesis of structurally diverse pleuromutilin antibiotics. Nat. Chem. 14, 1270–1277 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Cooper, C. J. et al. Molecular properties that define the activities of antibiotics in Escherichia coli and Pseudomonas aeruginosa. ACS Infect. Dis. 4, 1223–1234 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Mehla, J. et al. Predictive rules of efflux inhibition and avoidance in Pseudomonas aeruginosa. mBio 12, e02785–20 (2021).

  28. Leus Inga, V. et al. Functional diversity of Gram-negative permeability barriers reflected in antibacterial activities and intracellular accumulation of antibiotics. Antimicrob. Agents Chemother. 67, e01377–22 (2023).

  29. Geddes, E. J., Li, Z. & Hergenrother, P. J. An LC–MS/MS assay and complementary web-based tool to quantify and predict compound accumulation in E. coli. Nat. Protoc. 16, 4833–4854 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Wallace, M. J. et al. Discovery and characterization of the antimetabolite action of thioacetamide-linked 1,2,3-triazoles as disruptors of cysteine biosynthesis in Gram-negative bacteria. ACS Infect. Dis. 6, 467–478 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Vaara, M. Agents that increase the permeability of the outer membrane. Microbiol. Rev. 56, 395–411 (1992).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Huigens, R. W. 3rd et al. A ring-distortion strategy to construct stereochemically complex and structurally diverse compounds from natural products. Nat. Chem. 5, 195–202 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Perlmutter, S. J. et al. Compound uptake into E. coli can be facilitated by N-alkyl guanidiniums and pyridiniums. ACS Infect. Dis. 7, 162–173 (2021).

    Article  CAS  PubMed  Google Scholar 

  34. Hancock, R. E. & Woodruff, W. A. Roles of porin and beta-lactamase in beta-lactam resistance of Pseudomonas aeruginosa. Rev. Infect. Dis. 10, 770–775 (1988).

    Article  CAS  PubMed  Google Scholar 

  35. Ude, J. et al. Outer membrane permeability: antimicrobials and diverse nutrients bypass porins in Pseudomonas aeruginosa. Proc. Natl Acad. Sci. USA 118, e2107644118 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Loh, B., Grant, C. & Hancock, R. E. Use of the fluorescent probe 1-N-phenylnaphthylamine to study the interactions of aminoglycoside antibiotics with the outer membrane of Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 26, 546–551 (1984).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Hancock, R. E. & Farmer, S. W. Mechanism of uptake of deglucoteicoplanin amide derivatives across outer membranes of Escherichia coli and Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 37, 453–456 (1993).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Kung, V. L., Ozer, E. A. & Hauser, A. R. The accessory genome of Pseudomonas aeruginosa. Microbiol. Mol. Biol. Rev. 74, 621–641 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Mikkelsen, H., McMullan, R. & Filloux, A. The Pseudomonas aeruginosa reference strain PA14 displays increased virulence due to a mutation in ladS. PLoS ONE 6, e29113 (2011).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  40. Williams, J. J., Halvorsen, E. M., Dwyer, E. M., DiFazio, R. M. & Hergenrother, P. J. Toxin–antitoxin (TA) systems are prevalent and transcribed in clinical isolates of Pseudomonas aeruginosa and methicillin-resistant Staphylococcus aureus. FEMS Microbiol. Lett. 322, 41–50 (2011).

    Article  CAS  PubMed  Google Scholar 

  41. Surivet, J. P. et al. Synthesis and characterization of tetrahydropyran-based bacterial topoisomerase inhibitors with antibacterial activity against Gram-negative bacteria. J. Med. Chem. 60, 3776–3794 (2017).

    Article  CAS  PubMed  Google Scholar 

  42. Sum, P. E. et al. Glycylcyclines. 1. A new generation of potent antibacterial agents through modification of 9-aminotetracyclines. J. Med. Chem. 37, 184–188 (1994).

    Article  CAS  PubMed  Google Scholar 

  43. Smith, P. A. et al. Optimized arylomycins are a new class of Gram-negative antibiotics. Nature 561, 189–194 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

  44. Tanaka, N., Kinoshita, T. & Masukawa, H. Mechanism of protein synthesis inhibition by FA and related antibiotics. Biochem. Biophys. Res. Commun. 30, 278–283 (1968).

    Article  CAS  PubMed  Google Scholar 

  45. Garcia Chavez, M. et al. Synthesis of FA derivatives yields a potent antibiotic with an improved resistance profile. ACS Infect. Dis. 7, 493–505 (2021).

    Article  CAS  PubMed  Google Scholar 

  46. Haloi, N. et al. Rationalizing the generation of broad spectrum antibiotics with the addition of a positive charge. Chem. Sci. 12, 15028–15044 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Durand-Reville, T. F. et al. Rational design of a new antibiotic class for drug-resistant infections. Nature 597, 698–702 (2021).

    Article  ADS  CAS  PubMed  Google Scholar 

  48. Llanes, C. et al. Clinical strains of Pseudomonas aeruginosa overproducing MexAB-OprM and MexXY efflux pumps simultaneously. Antimicrob. Agents Chemother. 48, 1797–1802 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Skinner, S. O., Sepulveda, L. A., Xu, H. & Golding, I. Measuring mRNA copy number in individual Escherichia coli cells using single-molecule fluorescent in situ hybridization. Nat. Protoc. 8, 1100–1113 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

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

Download references

Acknowledgements

We thank the NIH (AI136773), the University of Illinois and Roche for support of this work. M.K.G. is a member of the NIH Chemistry-Biology Interface Training Grant (T32-GM136629). We thank L. Li (Metabolomics Center, Roy J. Carver Biotechnology Center, UIUC and Duke University School of Medicine Proteomics and Metabolomics Core Facility) for all LC–MS/MS analysis. We thank D. Olson, L. Zhu and N. Duay at the School of Chemical Sciences NMR Laboratory at UIUC for NMR services. We thank A. Cyphersmith and the Core Facilities at the Carl Woese Institute for Genomic Biology for assistance with confocal imaging. We are grateful to all past Hergenrother laboratory members who have contributed compounds to the complexity-to-diversity collection and to B. Drown, who wrote the code used for the random forest analysis. We thank H. Zgurskaya for the generous donation of the P. aeruginosa PA01 Δ6 and PA01 Δ6-pore strains and D. Bumann for the generous donation of the P. aeruginosa PA14 and PA14 Δ40 strain. We are further grateful to collaborators at Roche, including C. Kramer who helped establish this collaboration early on and G. M. Daniel for his guidance of the collaboration as alliance manager.

Author information

Authors and Affiliations

Authors

Contributions

P.J.H. and E.J.G. conceived the study. E.J.G. performed accumulation analyses, computational analyses and mode of uptake studies with assistance of M.K.G. E.J.G., M.R.L., M.K.G. and S.J.P. synthesized compounds in the test set. M.G.C., A.G. and M.K.G. synthesized FA derivatives. E.J.G., A.G., M.G.C. and M.K.G. performed MICs. C.B. and L.G. provided compounds and assisted with analysis of the data. E.J.G. and P.J.H. wrote this manuscript and it was approved by all authors. P.J.H. supervised this research.

Corresponding author

Correspondence to Paul J. Hergenrother.

Ethics declarations

Competing interests

The University of Illinois has filed patents on some of the compounds described herein on which P.J.H. and M.G.C. are inventors.

Peer review

Peer review information

Nature thanks Kim Lewis 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 figures and tables

Extended Data Fig. 1 Confocal fluorescent images of an intrinsically fluorescent compound, Bisindolylmaleimide X (BIM X).

a) The structure of BIM X along with accumulation values. The LC–MS/MS assay reveals BIM X as a poor accumulator in E. coli MG1655, while it is a good accumulator in P. aeruginosa PAO1. Accumulation units are in nmol/1012 CFU and the error is reported as the s.e.m. b) The standard accumulation assay was performed in P. aeruginosa (PAO1) and E. coli (MG1655) with either BIM X (50 μM) or DMSO until just after the oil removal step. Cells were fixed in 3.7 % formaldehyde in PBS and fluorescence images were taken with a Zeiss 710 multiphoton confocal microscope through a 63x/1.4 oil objective. All samples were excited at 488 nm with an argon laser and fluorescence emission was recorded from 599–689 nm. Images were acquired using Zen Black (Zen 2.3). DMSO controls for both cell types tested indicate no autofluorescence. Images are representative of n = 3 biologically independent samples.

Source Data

Extended Data Fig. 2 Comparison of species-specific accumulation trends.

a) In the expanded set of primary amines, many compounds that fit the eNTRy rules do not accumulate in P. aeruginosa PAO1 and many compounds that do not fit the eNTRy rules do accumulate in P. aeruginosa PAO1. Low globularity (glob) and low rotatable bonds (RB) are predictive for ~80% of compounds tested in E. coli MG1655, but only 41% of compounds in P. aeruginosa PAO1. Compounds with poor amine steric accessibility, low amphiphilic moment and multiple charges were removed from this analysis; 154 compounds total are included in the plots, see Compound Master Table for compounds. b) Compounds that accumulate in E. coli, but not in P. aeruginosa, can often be grouped according to structural class, but no additional trends were identified. c) Compounds that accumulate in P. aeruginosa, but not E. coli, are primarily compounds that have high rotatable bonds, high globularity, or both. Accumulation units are reported in nmol/1012 CFUs. n = 3 biologically independent samples. The s.e.m. is reported for accumulation values.

Source Data

Extended Data Fig. 3 Accumulation trends, matched molecular pairs and comparison of P. aeruginosa and E. coli accumulation.

a) Amphiphilic moment (vsurf_A) positively correlates with accumulation in both P. aeruginosa and E. coli. b) Amine steric accessibility is important for accumulation in both P. aeruginosa and E. coli. c) Similarly, compounds with primary amines on secondary carbons tend to accumulate higher than compounds with primary amines on tertiary carbons. d) Matched molecular pairs show a correlation between amine pKa and accumulation in P. aeruginosa. For structures, see Supplementary Table 8. The pKa estimation was calculated using the software MoKa (v3.2.2) from Molecular Discovery suite. e) Distribution of accumulation values in P. aeruginosa vs. E. coli for compounds that accumulate in both bacterial strains (131 compounds plotted, see Compound Master Table for structures). Accumulation levels were lower on average in P. aeruginosa PAO1 relative to E. coli MG1655. Accumulation units are reported in nmol/1012 CFUs. n = 3 biologically independent samples. The s.e.m. is reported for accumulation values. Strains used: E. coli MG1655, P. aeruginosa PAO1.

Source Data

Extended Data Fig. 4 Random forest prediction modelling.

a) Random forest prediction model results on dataset of all 240 primary amines, structures reported in Supplementary Table 2. ROC plot with 10 repeated cross-validations in the training classification models. b) Relative importance of top 15 descriptors for all 240 primary amines. c) Random forest prediction model results on dataset of 50 highest primary amine accumulators and 50 lowest primary amine accumulators (100 amines total; compounds 2.1–2.50; 2.191-2.240 in Supplementary Table 2). d) Relative importance of top 15 descriptors for 100 primary amines. e) Random forest prediction model results on dataset of 30 highest primary amine accumulators and 30 lowest primary amine accumulators (60 amines total; compounds 2.1–2.30; 2.211-2.240 in Supplementary Table 2). f) Relative importance of top 15 descriptors for 60 primary amines. Formal charge (h_pavgQ) and hydrogen bond donor surface area (vsa_don) are boxed in red. Early iterations of the model with the incomplete compound library always identified these properties within the top 15 most important.

Extended Data Fig. 5 Influence of magnesium on accumulation, evaluation of its binding by NMR and influence of magnesium and PMBN on compound accumulation.

a) Confocal fluorescent images of an intrinsically fluorescent compound, Bisindolylmaleimide X (BIM X) in P. aeruginosa with or without MgCl2. The standard accumulation assay was performed in P. aeruginosa (PAO1) with either DMSO, BIM X (50 μM), or BIM X (50 μM) + MgCl2 (1 mM), until just after the oil removal step. Cells were fixed in 3.7 % formaldehyde in PBS and fluorescence images were taken with a Zeiss 710 multiphoton confocal microscope through a 63x/1.4 oil objective. All samples were excited at 488 nm with an argon laser and fluorescence emission was recorded from 599–689 nm. Images were acquired using Zen Black (Zen 2.3). The images for P. aeruginosa + DMSO and P. aeruginosa + BIM X are the same as in Extended Data Fig. 1. DMSO controls indicate no autofluorescence. Images are representative of n = 3 biologically independent samples. b) Evaluation of accumulation of compounds in E. coli MG1655 upon cotreatment with MgCl2 (1 mM). c) Determination of Mg2+ interactions with polyamines. Ethylenediaminetetraacetic acid (EDTA), a compound known to chelate to Mg2+ in solution, clearly shows a marked chemical shift change in the NMR as well as changes in coupling when comparing EDTA alone to EDTA plus Mg2+. When this same experiment was performed with norspermine, there were no observable changes in coupling or chemical shift in the presence of MgCl2, suggesting a lack of perturbation of any electronic environment on the molecule, thus, a lack of interaction between Mg2+ and norspermine. Compounds were evaluated at a final concentration of 1 mM and MgCl2 was at a final concentration of 20 mM to maintain the same Mg2+-to-compound ratio as in the accumulation experiments. The pD of all solutions was adjusted to 7.2 prior to analysis. d) The same set of compounds as in Extended Data Fig. 5b showed a statistically significant increase in accumulation in P. aeruginosa PAO1 upon cotreatment with the permeabilizer PMBN (8 μg/ml). All structures and accumulation values are listed in Supplementary Table 4b. For samples with additives in Extended Data Fig. 5b & d, compounds 23 and 35 did not meet the mass spec standards and are thus excluded from this analysis. n = 3 biologically independent samples. The average and s.e.m are reported for accumulation values. Statistical significance was determined using a two-sample Welch’s t-test (one-tailed test, assuming unequal variance). Statistically significant accumulation differences for compounds in P. aeruginosa PA14 versus P. aeruginosa PA14 with PMBN treatment or E. coli MG1655 versus E. coli MG1655 with MgCl2 treatment are indicated with asterisks (n.s. not significant, *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001).

Source Data

Extended Data Fig. 6 Multiple amines and alternative positive charges often aid in accumulation in P. aeruginosa PAO1.

a) Diamines accumulate to significantly higher levels intracellularly in P. aeruginosa relative to their monoamine comparators. b) Diamines containing two primary amines consistently accumulate to a significant extent. c) Diamines containing one primary amine and one secondary or tertiary amine have more variable accumulation levels in P. aeruginosa, depending on hydrogen bond donor ability. d) Accumulation summary of 16 amine, guanidinium and pyridinium containing compounds (48 compounds total, structures in Supplementary Table 5) in P. aeruginosa PAO1. Compounds are classified as accumulators or non-accumulators based on statistical significance relative to the negative antibiotic controls. e) Examples of side-by-side amine (A), guanidinium (G) and pyridinium (P) comparators and their relative accumulation values in P. aeruginosa PAO1. Amine and guanidinium comparators tend to accumulate to a very similar extent, while pyridiniums often do not accumulate to a significant extent. f) Accumulation of antibiotic controls in three P. aeruginosa strains. Accumulation is consistent with antibacterial activity reported in Extended Data Table 1c.g) Accumulation of representative non-antibiotics in three P. aeruginosa strains. While there is some variance in accumulation levels between strains, high concordance of accumulation classification was observed. Compounds are classified as accumulators or non-accumulators based on statistical significance relative to the negative antibiotic controls. Structures and accumulation values are reported in Supplementary Table 6. Accumulation units are reported in nmol/1012 CFUs. n = 3 biologically independent samples. The average and s.e.m. are reported for accumulation values. Accumulation units are reported in nmol/1012 CFUs. All compounds were tested in biological triplicate. The average and s.e.m. are reported for accumulation values. Formal charge (FCH) was calculated using MOE.

Source Data

Extended Data Fig. 7 Accumulation and activity of FA and various derivatives and resistance generation in E. coli.

a) Introducing a hydrolyzable amidoxime linker onto FA provides a strategy to increase gram-negative activity and accumulation. Increasing the number of amines on the linker results in improved antibacterial activity and accumulation in two gram-negative species, E. coli MG1655 and P. aeruginosa PAO1, with the 4-amine linker-containing FA derivative (FA prodrug) demonstrating the most potent activity and highest absolute accumulation values. Accumulation values are reported in nmol/1012 CFUs and represent the concentration of free FA in the cell. n = 3 biologically independent samples. The average and s.e.m. are reported for accumulation values. b) E. coli MG1655 mutants resistant to FA prodrug were generated using a serial passage method and exhibit amino acid mutations mapping back to the FA binding pocket of EF-G. Bacteria (1 ×108 CFU/ml) were first inoculated in a standard MIC experiment against both sub- and supra-MIC antibiotic concentrations. The well containing the highest concentration of FA prodrug where bacterial growth was still observed was then isolated and re-subjected to this experiment. c) Acetylating both alcohols of FA greatly disrupts target engagement of FA with EF-G as inferred from MIC values against S. aureus. The diacetylated version of FA prodrug (FA prodrug Ac2O) loses 4–16x activity against both gram-positive and gram-negative bacterial strains, suggesting that inhibition of EF-G contributes to the observed antibacterial activity of FA prodrug. MICs were performed in MH or LB broth per Clinical and Laboratory Standards Institute (CLSI) guidelines. Accumulation values are reported in nmol/1012 CFUs and represent the concentration of free FA in the cell. n = 3 biologically independent samples. The average and s.e.m. are reported for accumulation values. * Indicates compound solubility limit.

Source Data

Extended Data Fig. 8 Mode of uptake and membrane interactions of FA prodrug.

a) (LEFT) In PBS at 37 °C, FA prodrug (10 µg/ml) hydrolyses to FA (FA) over a period of 48 h, as monitored via LC–MS. TIC LC–MS traces showing the disappearance of FA prodrug and the appearance of FA. Traces are all normalized to 2E6 ion counts. (RIGHT) Plasma stability, microsomal stability and 72-hour IC50 values for FA and FA prodrug. Average percentage errors are reported as s.e.m. IC50 errors reported as s.d. n = 3 biologically independent samples. b) Cotreatment with magnesium ions leads to a 32x increase in MIC for gentamicin and FA prodrug, while FA shows no change. Treatment with 80 mM NaCl shows minimal change when compared to the MgCl2 treated samples. MICs performed in P. aeruginosa PAO1 according to CLSI guidelines and values are reported in µg/ml. n = 3 biologically independent samples c) FA prodrug, gentamicin and colistin all permeabilize the outer-membrane of P. aeruginosa PAO1 to the membrane impermeable fluorophore NPN at a 10-minute time point, while treatment with FA shows no effect. Error bars represent the standard deviation from the average of RFUs. n = 3 biologically independent samples d) Treatment with FA prodrug leads to dose-dependent inner-membrane depolarization in P. aeruginosa PAO1, quantified using the potentiometric dye DiSC3 (5), while treatment with FA shows no effect. Concentrations of FA and FA prodrug are listed in µg/ml. 1% Triton X was used as the positive control, while 2% DMSO was used as the negative control. n = 3 biologically independent samples.

Source Data

Extended Data Table 1 MIC values of antibiotics against P. aeruginosa
Extended Data Table 2 Biological data on compounds included in the retrospective analysis comparing the properties of PA-active compounds versus less active derivatives

Supplementary information

Supplementary Information

Synthetic methods and NMR spectra for reported compounds.

Reporting Summary

Supplementary Tables

Supplementary Tables 1–8 and the Compound Master Table.

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

Geddes, E.J., Gugger, M.K., Garcia, A. et al. Porin-independent accumulation in Pseudomonas enables antibiotic discovery. Nature 624, 145–153 (2023). https://doi.org/10.1038/s41586-023-06760-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-023-06760-8

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

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

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