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 platform to investigate the functions of bacterial drug efflux pumps

Abstract

Efflux pumps are a serious challenge for the development of antibacterial agents. Overcoming efflux requires an in-depth understanding of efflux pump functions, specificities and the development of inhibitors. However, the complexities of efflux networks have limited such studies. To address these challenges, we generated Efflux KnockOut-35 (EKO-35), a highly susceptible Escherichia coli strain lacking 35 efflux pumps. We demonstrate the use of this strain by constructing an efflux platform comprising EKO-35 strains individually producing efflux pumps forming tripartite complexes with TolC. This platform was profiled against a curated diverse compound collection, which enabled us to define physicochemical properties that contribute to transport. We also show the E. coli drug efflux network is conditionally essential for growth, and that the platform can be used to investigate efflux pump inhibitor specificities and efflux pump interplay. We believe EKO-35 and the efflux platform will have widespread application for the study of drug efflux.

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: Phenotypic analysis of EKO-35 in nutrient-rich and -limited conditions.
Fig. 2: Characterizing the membrane-enriched proteome of EKO-35.
Fig. 3: The E. coli efflux system is contextually essential.
Fig. 4: EKO-35 and the efflux platform enabled summation of efflux substrate physicochemical properties.
Fig. 5: EKO-35 and the efflux platform can be used to assess EPI specificities and efflux pump interplay.

Similar content being viewed by others

Data availability

All supporting data are included in this published article and the Supplementary information. Source data are provided with this paper. EKO-35 and EKO-35-Pore genomic raw sequence data were deposited in GenBank (BioProject ID PRJNA838981). The reference E. coli genome was obtained from NCBI (accession no. CP009273.1). Mass spectrometry proteomics data have been deposited to the ProteomeXchange Consortium via the PRIDE partner repository with the dataset identifier PXD033975. Source data are provided with this paper.

References

  1. Tommasi, R., Brown, D. G., Walkup, G. K., Manchester, J. I. & Miller, A. A. ESKAPEing the labyrinth of antibacterial discovery. Nat. Rev. Drug Discov. 14, 529–542 (2015).

    CAS  PubMed  Google Scholar 

  2. Zhao, S. et al. Defining new chemical space for drug penetration into Gram-negative bacteria. Nat. Chem. Biol. 16, 1293–1302 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Piddock, L. J. V. Multidrug-resistance efflux pumps - not just for resistance. Nat. Rev. Microbiol. 4, 629–636 (2006).

    CAS  PubMed  Google Scholar 

  5. Teelucksingh, T., Thompson, L. K. & Cox, G. The evolutionary conservation of Escherichia coli drug efflux pumps supports physiological functions. J. Bacteriol. 202, e00367–20 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Saier, M. H. Jr. et al. Evolutionary origins of multidrug and drug-specific efflux pumps in bacteria. FASEB J. 12, 265–274 (1998).

    CAS  PubMed  Google Scholar 

  7. Nishino, K. & Yamaguchi, A. Analysis of a complete library of putative drug transporter genes in Escherichia coli. J. Bacteriol. 183, 5803–5812 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Saier, M. H. Jr. A functional-phylogenetic classification system for transmembrane solute transporters. Microbiol. Mol. Biol. Rev. 64, 354–411 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Sun, J., Deng, Z. & Yan, A. Bacterial multidrug efflux pumps: mechanisms, physiology and pharmacological exploitations. Biochem. Biophys. Res. Commun. 453, 254–267 (2014).

    CAS  PubMed  Google Scholar 

  10. Alcalde-Rico, M., Hernando-Amado, S., Blanco, P. & Martínez, J. L. Multidrug efflux pumps at the crossroad between antibiotic resistance and bacterial virulence. Front. Microbiol. 7, 1483 (2016).

    PubMed  PubMed Central  Google Scholar 

  11. Weston, N., Sharma, P., Ricci, V. & Piddock, L. J. V. Regulation of the AcrAB-TolC efflux pump in Enterobacteriaceae. Res. Microbiol. 169, 425–431 (2018).

    CAS  PubMed  Google Scholar 

  12. Krishnamoorthy, G. et al. Breaking the permeability barrier of Escherichia coli by controlled hyperporination of the outer membrane. Antimicrob. Agents Chemother. 60, 7372–7381 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Datsenko, K. A. & Wanner, B. L. One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. Proc. Natl Acad. Sci. USA 97, 6640–6645 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Jiang, Y. et al. Multigene editing in the Escherichia coli genome via the CRISPR-Cas9 system. Appl. Environ. Microbiol. 81, 2506–2514 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Baba, T. et al. Construction of Escherichia coli K-12 in-frame, single-gene knockout mutants: the Keio collection. Mol. Syst. Biol. https://doi.org/10.1038/msb4100050 (2006).

  16. Krin, E., Danchin, A. & Soutourina, O. Decrypting the H-NS-dependent regulatory cascade of acid stress resistance in Escherichia coli. BMC Microbiol. 10, 273 (2010).

    PubMed  PubMed Central  Google Scholar 

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

    Google Scholar 

  18. Jumper, J. et al. Highly accurate protein structure prediction with AlphaFold. Nature 596, 583–589 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Rest, M. E., van der, van der Rest, M. E., Frank, C. & Molenaar, D. Functions of the membrane-associated and cytoplasmic malate dehydrogenases in the citric acid cycle of Escherichia coli. J. Bacteriol. 182, 6892–6899 (2000).

    PubMed  PubMed Central  Google Scholar 

  20. Wang-Kan, X. et al. Lack of AcrB efflux function confers loss of virulence on Salmonella enterica serovar Typhimurium. mBio. https://doi.org/10.1128/mBio.00968-17 (2017).

  21. Cox, G. et al. A common platform for antibiotic dereplication and adjuvant discovery. Cell Chem. Biol. 24, 98–109 (2017).

    CAS  PubMed  Google Scholar 

  22. Zgurskaya, H. I., Krishnamoorthy, G., Ntreh, A. & Lu, S. Mechanism and function of the outer membrane channel TolC in multidrug resistance and physiology of Enterobacteria. Front. Microbiol. 2, 189 (2011).

    PubMed  PubMed Central  Google Scholar 

  23. Kitagawa, M. et al. Complete set of ORF clones of Escherichia coli ASKA library (a complete set of E. coli K-12 ORF archive): unique resources for biological research. DNA Res. 12, 291–299 (2005).

    CAS  PubMed  Google Scholar 

  24. Grinter, R. et al. Protease-associated import systems are widespread in Gram-negative bacteria. PLoS Genet. 15, e1008435 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Horiyama, T. & Nishino, K. AcrB, AcrD, and MdtABC multidrug efflux systems are involved in enterobactin export in Escherichia coli. PLoS ONE 9, e108642 (2014).

    PubMed  PubMed Central  Google Scholar 

  26. Hews, C. L., Cho, T., Rowley, G. & Raivio, T. L. Maintaining integrity under stress: envelope stress response regulation of pathogenesis in Gram-negative bacteria. Front. Cell. Infect. Microbiol. 9, 313 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Blanco, P. et al. Bacterial multidrug efflux pumps: much more than antibiotic resistance determinants. Microorganisms 4, 14 (2016).

  28. Deininger, K. N. W. et al. A requirement of TolC and MDR efflux pumps for acid adaptation and GadAB induction in Escherichia coli. PLoS ONE 6, e18960 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Lewinson, O., Padan, E. & Bibi, E. Alkalitolerance: a biological function for a multidrug transporter in pH homeostasis. Proc. Natl Acad. Sci. USA 101, 14073–14078 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Holdsworth, S. R. & Law, C. J. Multidrug resistance protein MdtM adds to the repertoire of antiporters involved in alkaline pH homeostasis in Escherichia coli. BMC Microbiol. 13, 113 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Zhang, Y. et al. The multidrug efflux pump MdtEF protects against nitrosative damage during the anaerobic respiration in Escherichia coli. J. Biol. Chem. 286, 26576–26584 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. El Zahed, S. S., French, S., Farha, M. A., Kumar, G. & Brown, E. D. Physicochemical and structural parameters contributing to the antibacterial activity and efflux susceptibility of small molecule inhibitors of Escherichia coli. Antimicrob. Agents Chemother. 65, e01925–20 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Smith, D. H. & Davis, B. D. Mode of action of novobiocin in Escherichia coli. J. Bacteriol. 93, 71–79 (1967).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Braun, M., Killmann, H., Maier, E., Benz, R. & Braun, V. Diffusion through channel derivatives of the Escherichia coli FhuA transport protein. Eur. J. Biochem. 269, 4948–4959 (2002).

    CAS  PubMed  Google Scholar 

  35. Srikumar, R., Kon, T., Gotoh, N. & Poole, K. Expression of Pseudomonas aeruginosa multidrug efflux pumps MexA-MexB-OprM and MexC-MexD-OprJ in a multidrug-sensitive Escherichia coli strain. Antimicrob. Agents Chemother. 42, 65–71 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Ramaswamy, V. K., Vargiu, A. V., Malloci, G., Dreier, J. & Ruggerone, P. Molecular rationale behind the differential substrate specificity of bacterial RND multi-drug transporters. Sci. Rep. 7, 8075 (2017).

    PubMed  PubMed Central  Google Scholar 

  37. Rosenberg, E. Y., Ma, D. & Nikaido, H. AcrD of Escherichia coli is an aminoglycoside efflux pump. J. Bacteriol. 182, 1754–1756 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Sulavik, M. C. et al. Antibiotic susceptibility profiles of Escherichia coli strains lacking multidrug dfflux pump genes. Antimicrob. Agents Chemother. 45, 1126–1136 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Lomovskaya, O. et al. Identification and characterization of inhibitors of multidrug resistance efflux pumps in Pseudomonas aeruginosa: novel agents for combination therapy. Antimicrob. Agents Chemother. 45, 105–116 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Anes, J., Sivasankaran, S. K., Muthappa, D. M., Fanning, S. & Srikumar, S. Exposure to sub-inhibitory concentrations of the chemosensitizer 1-(1-naphthylmethyl)-piperazine creates membrane destabilization in multi-drug resistant Klebsiella pneumoniae. Front. Microbiol. 10, 92 (2019).

    PubMed  PubMed Central  Google Scholar 

  41. Misra, R., Morrison, K. D., Cho, H. J. & Khuu, T. Importance of real-time assays to distinguish multidrug efflux pump-inhibiting and outer membrane-destabilizing activities in Escherichia coli. J. Bacteriol. 197, 2479–2488 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Lee, A. et al. Interplay between efflux pumps may provide either additive or multiplicative effects on drug resistance. J. Bacteriol. 182, 3142–3150 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Tal, N. & Schuldiner, S. A coordinated network of transporters with overlapping specificities provides a robust survival strategy. Proc. Natl Acad. Sci. USA 106, 9051–9056 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Vega, D. E. & Young, K. D. Accumulation of periplasmic enterobactin impairs the growth and morphology of Escherichia coli tolC mutants. Mol. Microbiol. 91, 508–521 (2014).

    CAS  PubMed  Google Scholar 

  45. Zwama, M., Yamaguchi, A. & Nishino, K. Phylogenetic and functional characterisation of the Haemophilus influenzae multidrug efflux pump AcrB. Commun. Biol. 2, 340 (2019).

    PubMed  PubMed Central  Google Scholar 

  46. Brown, D. G., May-Dracka, T. L., Gagnon, M. M. & Tommasi, R. Trends and exceptions of physical properties on antibacterial activity for Gram-positive and Gram-negative pathogens. J. Med. Chem. 57, 10144–10161 (2014).

    CAS  PubMed  Google Scholar 

  47. Takatsuka, Y., Chen, C. & Nikaido, H. Mechanism of recognition of compounds of diverse structures by the multidrug efflux pump AcrB of Escherichia coli. Proc. Natl Acad. Sci. USA 107, 6559–6565 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Johnson, J. W. et al. Antibacterial activity of metergoline analogues: revisiting the ergot alkaloid scaffold for antibiotic discovery. ACS Med. Chem. Lett. 13, 284–291 (2022).

    CAS  PubMed  Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

  50. Patel, J. B. et al. M07-A10: Methods for Dilution Antimicrobial Susceptibility Tests for Bacteria That Grow Aerobically; Approved Standard 10th edn (Clinical Laboratory Standards Institute, 2015).

  51. Hazel, A. J. et al. Conformational dynamics of AcrA govern multidrug efflux pump assembly. ACS Infect. Dis. 5, 1926–1935 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Rappsilber, J., Mann, M. & Ishihama, Y. Protocol for micro-purification, enrichment, pre-fractionation and storage of peptides for proteomics using StageTips. Nat. Protoc. 2, 1896–1906 (2007).

    CAS  PubMed  Google Scholar 

  53. Cox, J. & Mann, M. MaxQuant enables high peptide identification rates, individualized p.p.b.-range mass accuracies and proteome-wide protein quantification. Nat. Biotechnol. 26, 1367–1372 (2008).

    CAS  PubMed  Google Scholar 

  54. Cox, J. et al. Accurate proteome-wide label-free quantification by delayed normalization and maximal peptide ratio extraction, termed MaxLFQ. Mol. Cell. Proteom. 13, 2513–2526 (2014).

    CAS  Google Scholar 

  55. Tyanova, S. et al. The Perseus computational platform for comprehensive analysis of (prote)omics data. Nat. Methods 13, 731–740 (2016).

    CAS  PubMed  Google Scholar 

  56. Cox, J. & Mann, M. 1D and 2D annotation enrichment: a statistical method integrating quantitative proteomics with complementary high-throughput data. BMC Bioinf. 13, S12 (2012).

    CAS  Google Scholar 

Download references

Acknowledgements

This study was supported by a Natural Sciences and Engineering Research Council of Canada (NSERC) Discovery grant (no. RGPIN-2019-04996) and a Canada Foundation for Innovation (CFI) grant no. JELF 37730 awarded to G.C., a Postgraduate Scholarship (NSERC) awarded to L.K.T., an Ontario Graduate Scholarship awarded to S.Z. and a Canadian Graduate Scholarship awarded to J.A.G. This research was also supported by a Foundation grant from the Canadian Institutes of Health Research (grant no. FRN 143215), infrastructure funding from the CFI, a Research Excellence allocation from the Ontario Research Fund, and a Tier I Canada Research Chair award to E.D.B. We thank K. Klaus and H. Green-Glass for their assistance with aspects of efflux gene inactivation. We also thank T. Bhando for providing the E. coli Pore strain used for the generation of the EKO-35-Pore, H. Zgurskaya for donating the anti-AcrB antibody used in western blotting and C. Murphy and L. Carfrae for assistance with high-throughput susceptibility testing at the Center for Microbial Chemical Biology (CMCB), McMaster University. Last, we thank J. Krieger (Bioinformatics Solutions Inc.) for operating the mass spectrometer for the proteomics experiments and E. Roach (Advanced Analysis Center, University of Guelph) for technical assistance with generation of the scanning electron microscopy images.

Author information

Authors and Affiliations

Authors

Contributions

G.C. conceived the research. G.C. and E.D.B. guided the research. G.C., T.T., L.K.T. and C.R.M. designed the experiments. G.C., T.T., L.K.T., S.Z., S.E.G., C.R.M., N.M.K. and J.A.G. performed experiments. L.K.T. and S.Z. analyzed the genome sequence data. L.K.T. prepared the proteomic samples and analyzed the data with guidance from J.G.-M. The high-throughput susceptibility testing was performed by T.T., S.Z., N.M.K., C.R.M. and S.E.G. T.T. and S.Z. analyzed the data. G.C., T.T., L.K.T. and S.Z. wrote the manuscript and all authors commented on and approved the paper.

Corresponding author

Correspondence to Georgina Cox.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Chemical Biology thanks Jessica Blair, Karl Hassan 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 Plasmid-based complementation of genes harboring nonsynonymous mutations in EKO-35 is detrimental in EKO-35 and wild-type E. coli.

Growth profiling of EKO-35 with ASKA constructs expressing a, pitA b, rspA c, tufA d, yjfC e, wcaC f, gyrB in i, nutrient-rich Lysogeny broth and ii, M9 minimal glucose medium. Solid symbols represent empty vector controls while open symbols represent ASKA expressing constructs. Gene expression was induced with 0.1 mM IPTG and plasmid selection for EKO-35 was achieved using 1 µg/mL in M9 minimal glucose medium and 4 µg/mL in Lysogeny broth. Data points represent mean values of n = 3 biological replicates.

Source data

Extended Data Fig. 2 The growth of EKO-35 at acidic and alkaline pH is not restored through plasmid-based complementation of genes in EKO-35 harboring nonsynonymous genomic mutations.

EKO-35 growth is significantly reduced at a, pH 5.0 and is not restored through expression of rspA, pitA, tufa, gyrB, wcaC, and yjfC (P = 6.53 × 10−8, 5.34 × 10−7, 2.52 × 10−8, 3.16 × 10-7, 6.75 × 10−8, 4.80 × 10-8 respectively). Likewise, growth was not restored at b, pH 5.5 (P = 1.00 × 10−7, 1.03 × 10−7, 1.18 × 10−7, 1.45 × 10−6, 4.05 × 10−7, 4.02 × 10−4 respectively). EKO-35 expressing ASKA constructs did not grow at c, 8.5 and d, 9.0. Gene expression was induced with 0.1 mM IPTG and plasmid selection for EKO-35 was achieved using 1 µg/mL chloramphenicol. Data points represent mean values ± s.d. of n = 3 biological end-point readings after 18 h of growth. P-values were calculated by a two-tailed Student’s t-test (****P < 0.0001, ns = not significant) relative to EKO-35 pCA24N.

Source data

Extended Data Fig. 3 Plasmid-based complementation of genes in EKO-35 harboring genomic mutations alters biofilm formation in EKO-35 and wild-type E. coli.

Expression of rspA significantly increased biofilm formation in a, wild-type E. coli (P = 2.67 × 10−3) and b, EKO-35 (P = 3.35 × 10−2) in Lysogeny broth, whilst expression of pitA, tufA, gyrB, and yjfC significantly lowered biofilm formation in wild-type E. coli (P = 4.70 × 10−4, 2.98 × 10−5, 6.56 × 10−3, 1.09 × 10−3 respectively) and pitA, tufA, and yjfC in EKO-35 (P = 3.00 × 10−3, 1.65 × 10−1, 8.36 × 10−2, 3.09 × 10−3 respectively). Gene expression was induced using 0.1 mM IPTG and plasmid selection for EKO-35 was achieved using 1 µg/mL chloramphenicol in M9 minimal glucose medium and 4 µg/mL in Lysogeny broth. Data points represent mean values ± s.d. of n = 3 biological end-point readings after 24 h and 48 h of growth in nutrient-rich and -limited media, respectively. P-values were calculated by a two-tailed Student’s t-test (*P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001, ns = not significant).

Source data

Extended Data Fig. 4

Plasmid-based complementation of genes harboring nonsynonymous mutations in EKO-35 is detrimental in EKO-35 and wild-type E. coli under low oxygen conditions. Growth profiling of EKO-35 in low oxygen environments with ASKA constructs expressing a, pitA b, rspA c, tufA d, yjfC e, wcaC f, gyrB in i, Lysogeny broth with 10 mM KNO3 (1% oxygen) and ii, M9 minimal glucose medium (5% oxygen). Solid symbols represent empty vector controls while open symbols represent ASKA expressing constructs. Gene expression was induced with 0.1 mM IPTG and plasmid maintenance for EKO-35 was achieved using 1 µg/mL chloramphenicol in M9 minimal glucose medium and 4 µg/mL chloramphenicol in Lysogeny broth. Data points represent mean values of n = 3 biological replicates. P-values were calculated by a two-tailed Student’s t-test (*P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001, ns = not significant).

Source data

Extended Data Fig. 5 Assessing the susceptibility of EKO-35 harboring ASKA plasmids expressing rspA, tufA, pitA, wcaC, gyrB, and yjfC.

Strains were profiled against a, benzalkonium chloride, b, novobiocin, c, trimethoprim, d, synthetic 2, e, synthetic 14, f, synthetic 19. MICs were determined i, without and ii, with chloramphenicol to maintain ASKA plasmids. Gene expression was induced with 0.1 mM IPTG. 25 µg/mL and 4 µg/mL chloramphenicol were used to maintain ASKA plasmids in wild-type E. coli and EKO-35, respectively. Strains were assessed in technical duplicate.

Source data

Extended Data Fig. 6 Susceptibility and growth profiling of porinated wild-type (WT) K-12, ∆tolC, and EKO-35 strains in LB at 37 °C.

a, Measurement of growth kinetics revealed the pore marginally extended the lag phase of EKO-35-Pore (P = 0.013), and did not significantly impact the generation time of EKO-35-Pore (P = 0.29) compared to the non-porinated parental strains, which was assessed using n = 3 biological replicates. Statistical analysis was performed using a two-tailed Student’s t-test. b, Heatmap depicting vancomycin susceptibility of the porinated strains. Each strain was tested in technical duplicate and MIC values were normalized to 100%, where green represents the highest MIC value, and white represents the lowest value (see key). Susceptibility of the WT K-12, ∆tolC, and EKO-35 strains + /- the pore was assessed for 52 compounds. The physicochemical properties for each compound that resulted in a 4-fold or greater increase in susceptibility compared to the wild-type strain are presented as individual data points in the box plots. The line through the center of each box indicates the median, and whiskers the minimum and maximum values. Physicochemical properties were calculated using DataWarrior (Version 5.5.0): c, Molecular weight (MW); d, Lipophilicity (logP); e, Aqueous solubility (logS); and f, Polar Surface Area (PSA). n for each strain is defined as the number of active compounds that decreased the MIC value by ≥4-fold compared to the WT K-12 strain as listed in panel g. The bounds of the boxes represent 25% (Q1) to 75% (Q3) of the physicochemical substrate ranked ranges, which are also summarized in panel g. g, Summary of the physicochemical property ranges for each strain. Medians are indicated in parentheses.

Source data

Extended Data Fig. 7 Heat maps depicting susceptibility levels of strains expressing efflux pump-encoding genes.

Each strain was tested in technical duplicate and MIC values were normalized to 100% for each compound tested, where orange represents the highest MIC value, and white represents the lowest value (see key). Genes encoding a, AcrEF, b, AcrD, c, AcrB, d, MdtEF, e, MdtBC, f, MacAB, g, MacAB, and h, MexCD were chromosomally integrated into the wild-type strain, a single gene deletion mutant, and EKO-35. Susceptibility testing revealed no changes in the resistance levels of the wild-type or single gene deletion backgrounds, except for c, ΔacrB. EKO-35 was highly susceptible to all compounds tested, revealing susceptibility profiles similar to ΔtolC, and integration of efflux genes into this strain conferred resistance to the known substrates.

Source data

Extended Data Fig. 8 Heat maps depicting aminoglycoside susceptibility of the wild-type K-12 strain and an ΔacrD mutant expressing the chromosomally integrated acrD gene (araC::acrD).

Susceptibility testing was performed in both MHB and LB. a, Kanamycin in cation-adjusted Mueller Hinton II Broth (MHB II), b, Kanamycin in Lysogeny broth (LB), c, Gentamicin in MHB II, and d, Gentamicin in LB. The cell inoculum used was 104 cells/mL to replicate susceptibility testing conditions used in previous studies. Each strain was assessed using three technical replicates and MIC values were normalized to 100% for each compound tested, where orange represents the highest MIC value, and white represents the lowest value (see key).

Source data

Extended Data Fig. 9 EKO-35 and the efflux platform can be used to assess efflux pump inhibitor specificities.

Bar charts depicting fractional inhibitory concentration index (FICI) values of a, phenylalanine-arginine-β-naphthylamide (PAβN) in combination with fusidic acid and ciprofloxacin, in addition to FICI values of b, 1-(1-naphthylmethyl)-piperazine (NMP) in combination with fusidic acid and ciprofloxacin. The FICI represents the ΣFIC of each drug. The FIC for each drug was determined by dividing the MIC of each drug in combination, by the MIC of each drug alone. ΣFIC = FICA + FICB = (CA/MICA) + (CB/MICB). Synergy (FICI < 0.5), additive (FICI > 0.5–1.0), indifferent (FICI > 1.0–2.0), and antagonistic (FICI > 2.0), as highlighted on each bar chart. Related to Supplementary Datasets 3 and 4.

Source data

Extended Data Fig. 10 EKO-35 and the efflux platform can be used to assess efflux pump interplay.

Interplay was not observed for novobiocin in a, EKO-35 and b, EKO-35-Pore or for minocycline in c, EKO-35 and d, EKO-35-Pore. pGDP-2 harboring emrE is denoted as pEmrE. All integrated strains were transformed with the empty vector (pGDP-2). Genes encoding AcrB, AcrEF, AcrD, and MdtEF were integrated into the arabinose operon (araC) of EKO-35. Related to Fig. 5 and Supplementary Dataset 5. Data points represent mean MIC values for which the OD600nm value >0.100 ± s.d. of n = 3 biological replicates. P-values were calculated by a two-tailed Student’s t-test (ns = not significant).

Source data

Supplementary information

Supplementary Information

Supplementary Figs. 1–5 and Tables 1–13.

Reporting Summary

Supplementary Data 1

MICs of various compounds against the wild-type K-12, ΔtolC, EKO-35 and the efflux-integrated EKO-35 strains, which were used to calculate fold change.

Supplementary Data 2

MICs of various compounds against the wild-type (WT)-Pore, ΔtolC-Pore, EKO-35-Pore and the efflux-integrated EKO-35-Pore strains, which were used to calculate fold change.

Supplementary Data 3

The FICI for PAβN in combination with different antibiotics using EKO-35 and the efflux platform.

Supplementary Data 4

The FICI for NMP in combination with different antibiotics using EKO-35 and the efflux platform.

Supplementary Data 5

Assessing efflux pump interplay using EKO-35 and the efflux platform.

Supplementary Data 6

Strains and plasmids used in this study.

Supplementary Data 7

Primers and oligonucleotides used in this study.

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 1

Unprocessed western blots.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 9

Statistical source data.

Source Data Extended Data Fig. 10

Statistical source data.

Rights and permissions

Springer Nature or its licensor 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

Teelucksingh, T., Thompson, L.K., Zhu, S. et al. A genetic platform to investigate the functions of bacterial drug efflux pumps. Nat Chem Biol 18, 1399–1409 (2022). https://doi.org/10.1038/s41589-022-01119-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41589-022-01119-y

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