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:

Inhibiting fatty acid synthesis overcomes colistin resistance

Abstract

Treating multidrug-resistant infections has increasingly relied on last-resort antibiotics, including polymyxins, for example colistin. As polymyxins are given routinely, the prevalence of their resistance is on the rise and increases mortality rates of sepsis patients. The global dissemination of plasmid-borne colistin resistance, driven by the emergence of mcr-1, threatens to diminish the therapeutic utility of polymyxins from an already shrinking antibiotic arsenal. Restoring sensitivity to polymyxins using combination therapy with sensitizing drugs is a promising approach to reviving its clinical utility. Here we describe the ability of the biotin biosynthesis inhibitor, MAC13772, to synergize with colistin exclusively against colistin-resistant bacteria. MAC13772 indirectly disrupts fatty acid synthesis (FAS) and restores sensitivity to the last-resort antibiotic, colistin. Accordingly, we found that combinations of colistin and other FAS inhibitors, cerulenin, triclosan and Debio1452-NH3, had broad potential against both chromosomal and plasmid-mediated colistin resistance in chequerboard and lysis assays. Furthermore, combination therapy with colistin and the clinically relevant FabI inhibitor, Debio1452-NH3, showed efficacy against mcr-1 positive Klebsiella pneumoniae and colistin-resistant Escherichia coli systemic infections in mice. Using chemical genomics, lipidomics and transcriptomics, we explored the mechanism of the interaction. We propose that inhibiting FAS restores colistin sensitivity by depleting lipid synthesis, leading to changes in phospholipid composition. In all, this work reveals a surprising link between FAS and colistin resistance.

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: Biotin and FAS inhibitors reverse colistin resistance against mcr-1-expressing Enterobacteriaceae.
Fig. 2: Established mechanisms do not explain the synergy between colistin and MAC13772.
Fig. 3: Inhibiting biotin or FAS induces cell envelope stress by modifying phospholipid composition.
Fig. 4: In vivo efficacy of Debio1452-NH3 and colistin in a murine infection model.
Fig. 5: MAC13772 modifies the lipidome to increase membrane fluidity, overcoming colistin resistance in mcr-1-expressing E. coli.

Similar content being viewed by others

Data availability

RNA sequencing data are available at the NCBI Sequence Read Archive under accession PRJNA824525. Lipidomics data have been deposited to the EMBL-EBI MetaboLights database with the identifier MTBLS7286. The complete dataset can be accessed at https://www.ebi.ac.uk/metabolights/MTBLS7286. The suppressor genome sequences and reference strain E. coli C0279 are deposited in NCBI Sequence Read Archive, SRA: PRJNA889379. E. coli K-12 (U00096) reference genome was used in this study. The source data underlying Figs. 1c, 3a,b,g and 4c are provided as Supplementary Tables 1, 8 and 9. The data supporting the findings of this study are available from the corresponding author upon reasonable request. Source data are provided with this paper.

References

  1. Nation, R. L. & Li, J. Colistin in the 21st century. Curr. Opin. Infect. Dis. 22, 535–543 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Liu, Y. et al. Emergence of plasmid-mediated colistin resistance mechanism MCR-1 in animals and human beings in China: a microbiological and molecular biological study. Lancet Infect. Dis. 16, 161–168 (2016).

    Article  PubMed  Google Scholar 

  3. Li, Z., Cao, Y., Yi, L., Liu, J. & Yang, Q. Emergent polymyxin resistance: end of an era? Open Forum Infect. Dis. 6, ofz368 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  4. Manioglu, S. et al. Antibiotic polymyxin arranges lipopolysaccharide into crystalline structures to solidify the bacterial membrane. Nat. Commun. 13, 6195 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Sabnis, A. et al. Colistin kills bacteria by targeting lipopolysaccharide in the cytoplasmic membrane. eLife 10, e65836 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Poirel, L., Jayol, A. & Nordmann, P. Polymyxins: antibacterial activity, susceptibility testing, and resistance mechanisms encoded by plasmids or chromosomes. Clin. Microbiol. Rev. 30, 557–596 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Simpson, B. W. & Trent, M. S. Pushing the envelope: LPS modifications and their consequences. Nat. Rev. Microbiol. 17, 403–416 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. MacNair, C. R. et al. Overcoming mcr-1 mediated colistin resistance with colistin in combination with other antibiotics. Nat. Commun. 9, 458 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  9. Farha, M. A. et al. Uncovering the hidden antibiotic potential of cannabis. ACS Infect. Dis. 6, 338–346 (2020).

    Article  CAS  PubMed  Google Scholar 

  10. Domalaon, R. et al. The anthelmintic drug niclosamide synergizes with colistin and reverses colistin resistance in Gram-negative bacilli. Antimicrob. Agents Chemother. 63, e02574-18 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Minrovic, B. M., Jung, D., Melander, R. J. & Melander, C. New class of adjuvants enables lower dosing of colistin against Acinetobacter baumannii. ACS Infect. Dis. 4, 1368–1376 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Tsai, C. N. et al. Targeting two-component systems uncovers a small-molecule inhibitor of Salmonella virulence. Cell Chem. Biol. 27, 793–805 (2020).

    Article  CAS  PubMed  Google Scholar 

  13. Tyers, M. & Wright, G. D. Drug combinations: a strategy to extend the life of antibiotics in the 21st century. Nat. Rev. Microbiol. 17, 141–155 (2019).

    Article  CAS  PubMed  Google Scholar 

  14. Zimmerman, S. M., Lafontaine, A. A. J., Herrera, C. M., Mclean, A. B. & Trent, M. S. A whole-cell screen identifies small bioactives that synergize with polymyxin and exhibit antimicrobial activities against multidrug-resistant bacteria. Antimicrob. Agents Chemother. 64, e01677-19 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Mattingly, A. E. et al. Screening an established natural product library identifies secondary metabolites that potentiate conventional antibiotics. ACS Infect. Dis. 6, 2629–2640 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. O’Rourke, A. et al. Mechanism-of-action classification of antibiotics by global transcriptome profiling. Antimicrob. Agents Chemother. 64, e01207–e01219 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  17. Hood, M. I., Becker, K. W., Roux, C. M., Dunman, P. M. & Skaar, E. P. Genetic determinants of intrinsic colistin tolerance in Acinetobacter baumannii. Infect. Immun. 81, 542–551 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Carfrae, L. A. et al. Mimicking the human environment in mice reveals that inhibiting biotin biosynthesis is effective against antibiotic-resistant pathogens. Nat. Microbiol. 5, 93–101 (2020).

    Article  CAS  PubMed  Google Scholar 

  19. Zlitni, S., Ferruccio, L. F. & Brown, E. D. Metabolic suppression identifies new antibacterial inhibitors under nutrient limitation. Nat. Chem. Biol. 9, 796–804 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Breakpoint Tables for Interpretation of MICs and Zone Diameters Version 11.0 (The European Committee on Antimicrobial Susceptibility Testing, 2021); http://www.eucast.org

  21. Yang, Q. et al. Balancing mcr-1 expression and bacterial survival is a delicate equilibrium between essential cellular defence mechanisms. Nat. Commun. 8, 2054 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  22. Clifton, L. A. et al. Effect of divalent cation removal on the structure of Gram-negative bacterial outer membrane models. Langmuir 31, 404–412 (2015).

    Article  CAS  PubMed  Google Scholar 

  23. Tong, L. Structure and function of biotin-dependent carboxylases. Cell. Mol. Life Sci. 70, 863–891 (2013).

    Article  CAS  PubMed  Google Scholar 

  24. Vance, D., Goldberg, I., Mitsuhashi, O. & Bloch, K. Inhibition of fatty acid synthetases by the antibiotic cerulenin. Biochem. Biophys. Res. Commun. 48, 649–656 (1972).

    Article  CAS  PubMed  Google Scholar 

  25. Heath, R. J. et al. Mechanism of triclosan inhibition of bacterial fatty acid synthesis. J. Biol. Chem. 274, 11110–11114 (1999).

    Article  CAS  PubMed  Google Scholar 

  26. Ramage, B. et al. Comprehensive arrayed transposon mutant library of Klebsiella pneumoniae outbreak strain KPNIH1. J. Bacteriol. 199, e00352-17 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Chen, C. C. & Feingold, D. S. Locus of divalent cation inhibition of the bactericidal action of polymyxin B. Antimicrob. Agents Chemother. 2, 331–335 (1972).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Stokes, J. M. et al. Pentamidine sensitizes Gram-negative pathogens to antibiotics and overcomes acquired colistin resistance. Nat. Microbiol. 2, 17028 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Farha, M. A., Verschoor, C. P., Bowdish, D. & Brown, E. D. Collapsing the proton motive force to identify synergistic combinations against Staphylococcus aureus. Chem. Biol. 20, 1168–1178 (2013).

    Article  CAS  PubMed  Google Scholar 

  31. Sekyere, J. O. & Amoako, D. G. Carbonyl cyanide m-chlorophenylhydrazine (CCCP) reverses resistance to colistin, but not to carbapenems and tigecycline in multidrug-resistant Enterobacteriaceae. Front. Microbiol. 8, 228 (2017).

    Google Scholar 

  32. Majdalani, N., Heck, M., Stout, V. & Gottesman, S. Role of RcsF in signaling to the Rcs phosphorelay pathway in Escherichia coli. J. Bacteriol. 187, 6770–6778 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Price, N. L. & Raivio, T. L. Characterization of the Cpx regulon in Escherichia coli strain MC4100. J. Bacteriol. 191, 1798–1815 (2009).

    Article  CAS  PubMed  Google Scholar 

  35. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Yao, J. & Rock, C. O. Exogenous fatty acid metabolism in bacteria. Biochimie 141, 30–39 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Yao, J. & Rock, C. O. How bacterial pathogens eat host lipids: implications for the development of fatty acid synthesis therapeutics. J. Biol. Chem. 290, 5940–5946 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Shinitzky, M. & Barenholz, Y. Fluidity parameters of lipid regions determined by fluorescence polarization. Biochim. Biophys. Acta 515, 367–394 (1978).

    Article  CAS  PubMed  Google Scholar 

  39. 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 

  40. Feldman, M. F. et al. A promising bioconjugate vaccine against hypervirulent Klebsiella pneumoniae. Proc. Natl Acad. Sci. USA 116, 18655–18663 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Spapen, H., Jacobs, R., Gorp Van, V., Troubleyn, J. & Honoré, P. M. Renal and neurological side effects of colistin in critically ill patients. Ann. Intensive Care 1, 14 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  42. Nair, A. B. & Jacob, S. A simple practice guide for dose conversion between animals and human. J. Basic Clin. Pharm. 7, 27–31 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  43. Mykytczuk, N. C. S., Trevors, J. T., Leduc, L. G. & Ferroni, G. D. Fluorescence polarization in studies of bacterial cytoplasmic membrane fluidity under environmental stress. Prog. Biophys. Mol. Biol. 95, 60–82 (2007).

    Article  CAS  PubMed  Google Scholar 

  44. Denich, T. J., Beaudette, L. A., Lee, H. & Trevors, J. T. Effect of selected environmental and physico-chemical factors on bacterial cytoplasmic membranes. J. Microbiol. Methods 52, 149–182 (2003).

    Article  CAS  PubMed  Google Scholar 

  45. Mileykovskaya, E. & Dowhan, W. The Cpx two-component signal transduction pathway is activated in Escherichia coli mutant strains lacking phosphatidylethanolamine. J. Bacteriol. 179, 1029–1034 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Purcell, A. B., Voss, B. J. & Trent, M. S. Diacylglycerol kinase A is essential for polymyxin resistance provided by EptA, MCR-1, and other lipid A phosphoethanolamine transferases. J. Bacteriol. 204, e0049821 (2022).

    Article  PubMed  Google Scholar 

  47. Lepak, A. J., Wang, W. & Andes, D. R. Pharmacodynamic evaluation of MRX-8, a novel polymyxin, in the neutropenic mouse thigh and lung infection models against Gram-negative pathogens. Antimicrob. Agents Chemother. 64, e01517–e01520 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Brown, P. et al. Design of next generation polymyxins with lower toxicity: the discovery of SPR206. ACS Infect. Dis. 5, 1645–1656 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Li, J. et al. Total and semisyntheses of polymyxin analogues with 2-Thr or 10-Thr modifications to decipher the structure-activity relationship and improve the antibacterial activity. J. Med. Chem. 64, 5746–5765 (2021).

    Article  CAS  PubMed  Google Scholar 

  50. Roberts, K. D. et al. A synthetic lipopeptide targeting top-priority multidrug-resistant Gram-negative pathogens. Nat. Commun. 13, 1625 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. 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 

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

    Article  CAS  PubMed  Google Scholar 

  53. Performance Standards for Antimicrobial Susceptibility Testing 31st edn (CLSI, 2021).

  54. Lambert, R. J. W. & Lambert, R. A model for the efficacy of combined inhibitors. J. Appl. Microbiol. 95, 734–743 (2003).

    Article  CAS  PubMed  Google Scholar 

  55. Mangat, C. S., Bharat, A., Gehrke, S. S. & Brown, E. D. Rank ordering plate data facilitates data visualization and normalization in high-throughput screening. J. Biomol. Screen. 19, 1314–1320 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  56. Altschul, S. F., Gish, W., Miller, W., Myers, E. W. & Lipman, D. J. Basic local alignment search tool. J. Mol. Biol. 215, 403–410 (1990).

    Article  CAS  PubMed  Google Scholar 

  57. Keseler, I. M. et al. The EcoCyc database: reflecting new knowledge about Escherichia coli K-12. Nucleic Acids Res. 45, 543–550 (2017).

    Article  Google Scholar 

  58. Bakker, E. P. & Mangerich, W. E. Interconversion of components of the bacterial proton motive force by electrogenic potassium transport. J. Bacteriol. 147, 820–826 (1981).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Yan, A., Guan, Z. & Raetz, C. R. H. An undecaprenyl phosphate-aminoarabinose flippase required for polymyxin resistance in Escherichia coli. J. Biol. Chem. 282, 36077–36089 (2007).

    Article  CAS  PubMed  Google Scholar 

  60. Galanos, C., Lüderitz, O. & Westphal, O. A new method for the extraction of R lipopolysaccharides. Eur. J. Biochem. 9, 245–249 (1969).

    Article  CAS  PubMed  Google Scholar 

  61. Li, H. & Durbin, R. Fast and accurate long-read alignment with Burrows-Wheeler transform. Bioinformatics 26, 589–595 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  62. Liao, Y., Smyth, G. K. & Shi, W. FeatureCounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics 30, 923–930 (2014).

    Article  CAS  PubMed  Google Scholar 

  63. Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet J. https://doi.org/10.14806/ej.17.1.200 (2011).

  64. Chomczynski, P. & Sacchi, N. Single-step method of RNA isolation by acid guanidinium thiocyanate-phenol-chloroform extraction. Anal. Biochem. 162, 156–159 (1987).

    Article  CAS  PubMed  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  66. Karp, P. D. et al. Pathway Tools version 23.0 update: software for pathway/genome informatics and systems biology. Brief. Bioinform. 22, 109–126 (2021).

    Article  PubMed  Google Scholar 

  67. Supek, F., Bošnjak, M., Škunca, N. & Šmuc, T. Revigo summarizes and visualizes long lists of gene ontology terms. PLoS ONE 6, e21800 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Mohamed, A., Molendijk, J. & Hill, M. M. Lipidr: a software tool for data mining and analysis of lipidomics datasets. J. Proteome Res. 19, 2890–2897 (2020).

    Article  CAS  PubMed  Google Scholar 

  69. Calvo-Villamañán, A. et al. On-target activity predictions enable improved CRISPR-dCas9 screens in bacteria. Nucleic Acids Res. 48, e64 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  70. Depardieu, F. & Bikard, D. Gene silencing with CRISPRi in bacteria and optimization of dCas9 expression levels. Methods 172, 61–75 (2020).

    Article  CAS  PubMed  Google Scholar 

  71. Deatherage, D. E. & Barrick, J. E. Identification of mutations in laboratory-evolved microbes from next-generation sequencing data using breseq. Methods Mol. Biol. 1151, 165–188 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank G. Wright from McMaster University for bacterial strains from the Institute for Infectious Disease Research clinical collection, E. Parker and P. Hergenrother from the University of Illinois at Urbana-Champaign for providing ample Debio1452-NH3 for animal studies, M. Mulvey from the University of Manitoba for providing the environmental mcr-1-positive E. coli isolates, R. Melano at Public Health Ontario for bacterial strains GB687 and C0064, and D. Bikard from the Institut Pasteur for the plasmid, pFD152. The authors also thank the Centre for Microbial Chemical Biology at McMaster University, specifically N. Henriquez for LC–MS/MS and S. McCusker for providing bacterial strains. This research was supported by a Tier 1 Canada Research Chair award, a Foundation grant from the Canadian Institutes of Health Research (CHIR; FRN 143215), and a grant from the Ontario Research Fund (RE09-047) to E.D.B. L.A.C. was supported by a CIHR Canada Graduate Scholarship (CGS-D) scholarship. The funders had no role in study design, data collection and analysis, decision to publish or preparation of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

L.A.C. conceived the research, designed and carried out experiments and data analysis and wrote the manuscript. K.R. assisted with the design and creation of CRISPRi strains. S.F. acquired samples for lipidomic analysis and S.F. with R.G. aided in the analysis of the lipidomics dataset. L.S., O.G.O. and B.R.C. performed lipid A purification and mass spec analysis. C.N.T. assisted in the analysis of the RNA-seq dataset. M.M.T. assisted in the animal experiments. C.R.M. acquired motility data and assisted with manuscript editing. C.W. assisted with data interpretation and manuscript editing. E.D.B. conceived the research and assisted with data interpretation and manuscript editing.

Corresponding author

Correspondence to Eric D. Brown.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Microbiology thanks Eefjan Breukink, Peter Tonge, Judith Steenbergen 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 Colistin synergizes with biotin and FAS inhibitors against E. coli strains harbouring mcr-1 on natural plasmids.

(a, b) Chequerboard broth microdilution assays showing dose-dependent colistin potentiation by cerulenin, MAC13772, and triclosan against mcr-1 positive E. coli strains (a) N15-02865 and (b) N15-02866. Dark regions represent higher cell density. Chequerboard data are representative of at least three biological replicates.

Source data

Extended Data Fig. 2 Exogenous LPS or Mg2+ abolishes synergy between MAC13772 and colistin against mcr-1 expressing E. coli.

Chequerboard broth microdilution assays between MAC13772 and colistin in M9 minimal media supplemented with (a) purified E. coli LPS (1 mg/mL) or (b) Mg2+ (20 mM). Dark regions represent higher cell density. Chequerboard data are representative of at least 2 biological replicates.

Source data

Extended Data Fig. 3 MAC13772 synergizes with colistin against E. coli expressing mcr-1 by inhibiting BioA.

(a,b) Chequerboard broth microdilution assays between MAC13772 and colistin M9 minimal media supplemented with biotin (10 µg/mL). (c) Expression of bioA suppresses MAC13772’s potentiation of colistin against mcr-1 positive E. coli. Expression of bioA induced by the addition of IPTG (10 mM). (d, e) Chequerboard assay of colistin and anhydrotetracycline against E. coli expressing (d) pfD152 empty or (e) pfD152 bioA. Dark regions represent higher cell density. Chequerboard data are representative of at least two biological replicates.

Source data

Extended Data Fig. 4 Synergy between MAC13772 and colistin is dependent on PEtN modification of lipid A.

Chequerboard broth microdilution assays between colistin and (a) MAC13772, (b) cerulenin, or (c) triclosan against E. coli expressing mcr-1::E246A. Dark regions represent higher cell density. Chequerboard data are representative of at least three biological replicates.

Source data

Extended Data Fig. 5 Synergy between MAC13772 and colistin requires L-Ara4N decoration of lipid A.

(a) Screening of the K. pneumoniae transposon mutant library for sensitivity to colistin. Percent growth was calculated based on growth in the presence of colistin relative to its respective control (untreated strain) growing in M9 minimal media. The blue region indicates genes 3 standard deviations from the mean. Bacterial pathways enriched among these mutants include genes involved in lipid A modification (arnA, arnB, arnC, arnD, arnT, phoPQ), cell division (envC, ftsK, ftsN, ftsX, nlpD, pal, tolA, tolB, tolQ, tolR), and enterobactin biosynthesis (entA, entD, entE, fes). (b) Chequerboard assays of MAC13772 and colistin against wild-type and colistin sensitive transposon mutants. Higher growth is indicated in dark blue, no detectable growth is indicated in white, and results are representative of at least two independent experiments.

Source data

Extended Data Fig. 6 Colistin does not increase intracellular accumulation of cerulenin and triclosan.

(a–c) Accumulation of (a) cerulenin (50 μM; blue, n = 4), (b) triclosan (50 μM; green, n = 5, 4), and (c) rifampicin (50 μM; red, n = 4; p = 0.0143) in mcr-1 expressing E. coli in the absence and presence of colistin (6 μM). Statistical significance was determined by using a one-tailed Mann–Whitney U test relative to the no colistin control, *P < 0.05.

Source data

Extended Data Fig. 7 Inhibiting biotin or FAS does not disrupt the proton motive force.

(a) DiSC3(5) assay, bar plots depict the mean of three biological replicates, error bars indicate standard error. (b, c) Chequerboard broth microdilution assays for (b) cerulenin or (c) triclosan and CCCP. Higher growth is indicated in dark blue; no detectable growth is indicated in white, and results are representative of at least two independent experiments.

Source data

Extended Data Fig. 8 Bacterial lysis assays with colistin and MAC13772 against empty and mcr-1 expressing E. coli.

(a) Killing of E. coli expressing the empty vector in the presence of water (n = 3), colistin (0.4 μg/mL; n = 4), and colistin (2 μg/mL; n = 5). (b, c) Killing of mcr-1 expressing E. coli in the presence of (b) water (n = 2), colistin (2 μg/mL; t = 4 and 8: n = 4, t = 0, 4, 24: n = 5), colistin (4 μg/mL; t = 0, 2, 4: n = 5, t = 8 and 24: n = 4), colistin (8 μg/mL; n = 2), colistin (32 μg/mL; n = 2) or (c) DMSO, MAC13772 (32 μg/mL; n = 4), MAC13772 (64 μg/mL; n = 2), MAC13772 (128 μg/mL; t = 0, 2, 4, 24: n = 2), MAC13772 (256 μg/mL; t = 0, 2, 4, 24: n = 2), MAC13772 (512 μg/mL; n = 2). The initial cell density is ~5 × 107 CFU/mL. Shown is the mean of a minimum of two biological replicates. Bars denote standard error.

Source data

Extended Data Fig. 9 Addition of exogenous fatty acids suppresses synergy between colistin and cerulenin or triclosan.

(a) Diagram depicting the path for the incorporation of exogenous and endogenous fatty acids in LPS and phospholipids. E. coli synthesizes phosphatidic acid (PA), the major precursor to all phospholipid species, through two acylation reactions. First glycerol-3-phosphate (G3P) is acylated to lysophosphatidic acid (LPA) by PlsB, which is subsequently acylated by PlsC to form phosphatidic acid. The acyltransferases (PlsB and PlsC) can use both acyl-CoA, generated by FadD from exogenous fatty acids, and acyl-ACP from FAS as acyl donors in the synthesis of phosphatidic acid. LPS biosynthesis requires endogenous β-hydroxyacyl-ACP for acylation reactions as there is not a pathway in E. coli to generate β-hydroxyacyl-ACP from exogenous fatty acids. (b–d) Chequerboard broth microdilution assays between cerulenin, MAC13772, and triclosan with colistin in media supplemented with (b) 2-hexadecenoic acid, (c) linoleic acid, or (d) oleic acid. Higher growth is indicated in dark blue, no detectable growth is indicated in white, and results are representative of at least three independent experiments.

Source data

Extended Data Fig. 10 Depletion of phosphatidylethanolamine biosynthesis genes sensitize mcr-1 expressing E. coli to colistin.

Chequerboard assay of colistin and anhydrotetracycline against E. coli expressing (a) mcr-1 or (b) empty vector and with a CRISPR knockdown of cdsA, pssA, psd, or pgsA (left to right). Dark regions represent higher cell density. Chequerboard data are representative of at least three biological replicates.

Source data

Supplementary information

Supplementary Information

Supplementary figs. 1–15, tables 2, 3, 5, 6 and 9–11, and unprocessed scans for supplementary figures.

Reporting Summary

Supplementary Table

Table 1. Interactions of cerulenin, MAC13772 and triclosan with colistin against a broad spectrum of bacteria. Minimum inhibitory concentration represents the median of at least two biological replicates. FICi were calculated with colistin using chequerboard assays and are the mean of at least two biological replicates. The fold change indicates the change in colistin MIC in the presence of ¼ MIC of partner compound and represents the median of at least two biological replicates (when an even number of replicates was performed, the mean of the two median values is shown). Short forms, Col, Cer, MAC and Tri indicate colistin, cerulenin, MAC13772 and triclosan. Results corresponding to the grey shaded squares were not assessed. Table 4. Screening data for K. pneumoniae MKP103 transposon library grown in M9 supplemented with amino acids and subinhibitory colistin. Normalized growth data associated with genetic screen in M9 and colistin. Mutants that exhibited growth <3 s.d. from the mean in the presence of colistin, but not M9 minimal media, were considered as exclusively defective for growth in the presence of colistin. Table 5. Gene ontology (GO) term enrichment for K. pneumoniae MKP103 transposon library grown in M9 supplemented with amino acids and subinhibitory colistin. The table depicts the top GO terms enriched by biological function. Enrichment was tested using pathway-tools software (Grossmann’s parent-child-union variation of the Fisher-exact test with the Benjamini-Hochberg correction). Table 8. Impact of mcr-1 expression and MAC13772 treatment on E. coli transcriptome. Raw values used to generate RNA sequencing heat maps and volcano plots. A Wald test corrected for multiple testing using the Benjamini and Hochberg method was used to calculate differentially expressed genes. Table 9. Impact of mcr-1 expression and MAC13772 treatment on E. coli lipidome. A moderated t statistic was used to identify significantly differential lipids between biological conditions applying a multiple hypothesis testing correction. Table 10. Identified mutations in an E. coli mutant resistant to MAC13772-colistin synergism.

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

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 (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

Carfrae, L.A., Rachwalski, K., French, S. et al. Inhibiting fatty acid synthesis overcomes colistin resistance. Nat Microbiol 8, 1026–1038 (2023). https://doi.org/10.1038/s41564-023-01369-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41564-023-01369-z

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research