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 Gram-negative-selective antibiotic that spares the gut microbiome

Abstract

Infections caused by Gram-negative pathogens are increasingly prevalent and are typically treated with broad-spectrum antibiotics, resulting in disruption of the gut microbiome and susceptibility to secondary infections1,2,3. There is a critical need for antibiotics that are selective both for Gram-negative bacteria over Gram-positive bacteria, as well as for pathogenic bacteria over commensal bacteria. Here we report the design and discovery of lolamicin, a Gram-negative-specific antibiotic targeting the lipoprotein transport system. Lolamicin has activity against a panel of more than 130 multidrug-resistant clinical isolates, shows efficacy in multiple mouse models of acute pneumonia and septicaemia infection, and spares the gut microbiome in mice, preventing secondary infection with Clostridioides difficile. The selective killing of pathogenic Gram-negative bacteria by lolamicin is a consequence of low sequence homology for the target in pathogenic bacteria versus commensals; this doubly selective strategy can be a blueprint for the development of other microbiome-sparing antibiotics.

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: Identification of lolamicin.
Fig. 2: Antimicrobial assessment and resistance frequency studies of lolamicin.
Fig. 3: Major binding and transient binding sites of lolamicin in LolCDE.
Fig. 4: In vivo efficacy studies with lolamicin.
Fig. 5: Lolamicin spares the gut microbiome and prevents C. difficile colonization.

Similar content being viewed by others

Data availability

All data supporting the findings of this study are available within the paper and the Supplementary Information. Raw sequencing data have been deposited at the NCBI Sequence Read Archive under accession PRJNA1101557. Source data are provided for Figs. 2, 4 and 5, Extended Data Figs. 14Source data are provided with this paper.

Code availability

Source code for data analysis can be found at https://github.com/HPCBio/hergenrother-16S-mouse-2022Sept and https://doi.org/10.5281/zenodo.10980655 (ref. 80).

References

  1. Maier, L. et al. Unravelling the collateral damage of antibiotics on gut bacteria. Nature 599, 120–124 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  2. Lynch, S. V. & Pedersen, O. The human intestinal microbiome in health and disease. New Eng. J. Med. 375, 2369–2379 (2016).

    Article  CAS  PubMed  Google Scholar 

  3. Schubert, A. M., Sinani, H. & Schloss, P. D. Antibiotic-induced alterations of the murine gut microbiota and subsequent effects on colonization resistance against Clostridium difficile. mBio 6, e00974 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Owens, R. C. Jr., Donskey, C. J., Gaynes, R. P., Loo, V. G. & Muto, C. A. Antimicrobial-associated risk factors for Clostridium difficile infection. Clin. Infect. Dis. 46, S19–S31 (2008).

    Article  PubMed  Google Scholar 

  5. Iizumi, T., Battaglia, T., Ruiz, V. & Perez Perez, G. I. Gut microbiome and antibiotics. Arch. Med. Res. 48, 727–734 (2017).

    Article  CAS  PubMed  Google Scholar 

  6. Poon, S. S. B. et al. Neonatal antibiotics have long term sex-dependent effects on the enteric nervous system. J. Phys. 600, 4303–4323 (2022).

    CAS  Google Scholar 

  7. Lange, K., Buerger, M., Stallmach, A. & Bruns, T. Effects of antibiotics on gut microbiota. Digest. Dis. 34, 260–268 (2016).

    Article  Google Scholar 

  8. Gu, S. et al. Effect of the short-term use of fluoroquinolone and β-lactam antibiotics on mouse gut microbiota. Infect. Drug Resist. 13, 4547–4558 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Lofmark, S., Jernberg, C., Jansson, J. K. & Edlund, C. Clindamycin-induced enrichment and long-term persistence of resistant Bacteroides spp. and resistance genes. J. Antimicrob. Chemother. 58, 1160–1167 (2006).

    Article  PubMed  Google Scholar 

  10. Hertz, F. B. et al. Effects of antibiotics on the intestinal microbiota of mice. Antibiotics 9, 191 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Lagier, J. C., Million, M., Hugon, P., Armougom, F. & Raoult, D. Human gut microbiota: repertoire and variations. Front. Cell Infect. Microbiol. 2, 136 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  12. Singh, H. Management with colistin. Ind. J. Crit. Care. Med. 14, 161–162 (2010).

    Article  Google Scholar 

  13. Falagas, M. E. & Kasiakou, S. K. Colistin: the revival of polymyxins for the management of multidrug-resistant Gram-negative bacterial infections. Clin. Infect. Dis. 40, 1333–1341 (2005).

    Article  CAS  PubMed  Google Scholar 

  14. Chatzidimitriou, M. et al. mcr genes conferring colistin resistance in Enterobacterales; a five year overview. Acta Med. Acad. 50, 365–371 (2021).

    Article  PubMed  Google Scholar 

  15. Rice, L. B. Federal funding for the study of antimicrobial resistance in nosocomial pathogens: no ESKAPE. J. Infect. Dis. 197, 1079–1081 (2008).

    Article  PubMed  Google Scholar 

  16. Hoffman, P. S. Antibacterial discovery: 21st century challenges. Antibiotics 9, 213–213 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  17. Nickerson, N. N. et al. A novel inhibitor of the LolCDE ABC transporter essential for lipoprotein trafficking in Gram-negative bacteria. Antimicrob. Agents Chemother. 62, e02151–02117 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Zhang, G. et al. Cell-based screen for discovering lipopolysaccharide biogenesis inhibitors. Proc. Natl Acad. Sci. USA 115, 6834–6839 (2018).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  19. Lehman, K. M. & Grabowicz, M. Countering Gram-negative antibiotic resistance: recent progress in disrupting the outer membrane with novel therapeutics. Antibiotics 8, 163 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Brown, M. F. et al. Potent inhibitors of LpxC for the treatment of Gram-negative infections. J. Med. Chem. 55, 914–923 (2012).

    Article  CAS  PubMed  Google Scholar 

  21. Miller, R. D. et al. A novel antibiotic targeting BamA identified by a computational search. Nat. Microbiol. 7, 1661–1672 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Imai, Y. et al. A new antibiotic selectively kills Gram-negative pathogens. Nature 576, 459–464 (2019).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  23. 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 

  24. Tokuda, H. & Matsuyama, S.-I. Sorting of lipoproteins to the outer membrane in E. coli. Biochim. Biophys. Acta 1693, 5–13 (2004).

    Article  CAS  PubMed  Google Scholar 

  25. Pathania, R. et al. Chemical genomics in Escherichia coli identifies an inhibitor of bacterial lipoprotein targeting. Nat. Chem. Biol. 5, 849–856 (2009).

    Article  CAS  PubMed  Google Scholar 

  26. Barker, C. A. et al. Degradation of MAC13243 and studies of the interaction of resulting thiourea compounds with the lipoprotein targeting chaperone LolA. Bioorg. Med. Chem. Lett. 23, 2426–2431 (2013).

    Article  CAS  PubMed  Google Scholar 

  27. Hoang, H. H. et al. Outer membrane targeting of Pseudomonas aeruginosa proteins shows variable dependence on the components of Bam and Lol machineries. mBio 2, e00246–00211 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Ito, H. et al. A new screening method to identify inhibitors of the Lol (localization of lipoproteins) system, a novel antibacterial target. Microbiol. Immun. 51, 263–270 (2007).

    Article  CAS  Google Scholar 

  29. Nayar, A. S. et al. Novel antibacterial targets and compounds revealed by a high-throughput cell wall reporter assay. J. Bacteriol. 197, 1726–1734 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Liu, J. et al. Natural inhibitors targeting the localization of lipoprotein system in Vibrio parahaemolyticus. Int. J. Mol. Sci. 23, 14352 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Breidenstein, E. B. M. et al. SMT-738: a novel small-molecule inhibitor of bacterial lipoprotein transport targeting Enterobacteriaceae. Antimicrob. Agents Chemother. 68, e0069523 (2024).

    Article  CAS  PubMed  Google Scholar 

  32. Grabowicz, M. & Silhavy, T. J. Redefining the essential trafficking pathway for outer membrane lipoproteins. Proc. Natl Acad. Sci. USA 114, 4769–4774 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  33. 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 

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

    Article  ADS  CAS  PubMed  Google Scholar 

  35. Munoz, K. A. & Hergenrother, P. J. Facilitating compound entry as a means to discover antibiotics for Gram-negative bacteria. Acc. Chem. Res. 54, 1322–1333 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Prochnow, H. et al. Subcellular quantification of uptake in Gram-negative bacteria. Anal. Chem. 91, 1863–1872 (2019).

    Article  CAS  PubMed  Google Scholar 

  37. Wexler, H. M. Bacteroides: the good, the bad, and the nitty-gritty. Clin. Microbiol. Rev. 20, 593–621 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Nonejuie, P., Burkart, M., Pogliano, K. & Pogliano, J. Bacterial cytological profiling rapidly identifies the cellular pathways targeted by antibacterial molecules. Proc. Natl Acad. Sci. USA 110, 16169–16174 (2013).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  39. Grabowicz, M. Lipoproteins and their trafficking to the outer membrane. EcoSal Plus 8, https://doi.org/10.1128/ecosalplus.ESP-0038-2018 (2019).

  40. Tang, X. et al. Structural basis for bacterial lipoprotein relocation by the transporter LolCDE. Nat. Struct. Mol. Biol. 28, 347–355 (2021).

    Article  CAS  PubMed  Google Scholar 

  41. Sharma, S. et al. Mechanism of LolCDE as a molecular extruder of bacterial triacylated lipoproteins. Nat. Commun. 12, 4687 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  42. Buffie, C. G. et al. Profound alterations of intestinal microbiota following a single dose of clindamycin results in sustained susceptibility to Clostridium difficile-induced colitis. Infect. Immun. 80, 62–73 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Lesniak, N. A., Schubert, A. M., Sinani, H. & Schloss, P. D. Clearance of Clostridioides difficile colonization is associated with antibiotic-specific bacterial changes. mSphere 6, e01238-20 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  44. Feuerstadt, P., Theriault, N. & Tillotson, G. The burden of CDI in the United States: a multifactorial challenge. BMC Infect. Dis. 23, 132 (2023).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  46. Oefner, C. et al. Increased hydrophobic interactions of iclaprim with Staphylococcus aureus dihydrofolate reductase are responsible for the increase in affinity and antibacterial activity. J. Antimicrob. Chemother. 63, 687–698 (2009).

    Article  CAS  PubMed  Google Scholar 

  47. Purnapatre, K. P. et al. In vitro and in vivo activities of DS86760016, a novel leucyl-tRNA synthetase inhibitor for Gram-negative pathogens. Antimicrob. Agents Chemother. 62, e01987-17 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  48. Schuster, M. et al. Peptidomimetic antibiotics disrupt the lipopolysaccharide transport bridge of drug-resistant Enterobacteriaceae. Sci. Adv. 9, eadg3683 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Rana, P. et al. FabI (enoyl acyl carrier protein reductase)—a potential broad spectrum therapeutic target and its inhibitors. Eur. J. Med. Chem. 208, 112757 (2020).

    Article  CAS  PubMed  Google Scholar 

  50. 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 

  51. Yao, J. et al. A pathogen-selective antibiotic minimizes disturbance to the microbiome. Antimicrob. Agents Chemother. 60, 4264–4273 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Konovalova, A., Kahne, D. E. & Silhavy, T. J. Outer membrane biogenesis. Annu. Rev. Microbiol. 71, 539–556 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Ho, H. et al. Structural basis for dual-mode inhibition of the ABC transporter MsbA. Nature 557, 196–201 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  55. Pandit, K. R. & Klauda, J. B. Membrane models of E. coli containing cyclic moieties in the aliphatic lipid chain. Biochim. Biophys. Acta 1818, 1205–1210 (2012).

    Article  CAS  PubMed  Google Scholar 

  56. Jorgensen, W. L., Chandrasekhar, J., Madura, J. D., Impey, R. W. & Klein, M. L. Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 79, 926–935 (1983).

    Article  ADS  CAS  Google Scholar 

  57. Jo, S., Kim, T., Iyer, V. G. & Im, W. CHARMM-GUI: a web-based graphical user interface for CHARMM. J. Comput. Chem. 29, 1859–1865 (2008).

    Article  CAS  PubMed  Google Scholar 

  58. Licari, G., Dehghani-Ghahnaviyeh, S. & Tajkhorshid, E. Membrane Mixer: a toolkit for efficient shuffling of lipids in heterogeneous biological membranes. J. Chem. Inform. Model. 62, 986–996 (2022).

    Article  CAS  Google Scholar 

  59. Humphrey, W., Dalke, A. & Schulten, K. VMD: visual molecular dynamics. J. Mol. Graph. 14, 27–38 (1996). 33-38.

    Article  Google Scholar 

  60. Trott, O. & Olson, A. J. AutoDock Vina: improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading. J. Comput. Chem. 31, 455–461 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Phillips, J. C. et al. Scalable molecular dynamics on CPU and GPU architectures with NAMD. J. Chem. Phys. 153, 044130 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Phillips, J. C. et al. Scalable molecular dynamics with NAMD. J. Comput. Chem. 26, 1781–1802 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Hart, K. et al. Optimization of the CHARMM additive force field for DNA: improved treatment of the BI/BII conformational equilibrium. J. Chem. Theor. Comput. 8, 348–362 (2012).

    Article  CAS  Google Scholar 

  64. Klauda, J. B. et al. Update of the CHARMM all-atom additive force field for lipids: validation on six lipid types. J. Phys. Chem. B 114, 7830–7843 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Darden, T., York, D. & Pedersen, L. Particle mesh Ewald: an N·log(N) method for Ewald sums in large systems. J. Chem. Phys. 98, 10089–10092 (1993).

    Article  ADS  CAS  Google Scholar 

  66. Ryckaert, J. P., Ciccotti, G. & Berendsen, H. J. C. Numerical integration of the Cartesian equations of motion of a system with constraints: molecular dynamics of n-alkanes. J. Comput. Phys. 23, 327–341 (1977).

    Article  ADS  CAS  Google Scholar 

  67. Martyna, G. J., Tobias, D. J. & Klein, M. L. Constant-pressure molecular-dynamics algorithms. J. Chem. Phys. 101, 4177–4189 (1994).

    Article  ADS  CAS  Google Scholar 

  68. Feller, S. E., Zhang, Y. H., Pastor, R. W. & Brooks, B. R. Constant-pressure molecular-dynamics simulation—the Langevin piston method. J. Chem. Phys. 103, 4613–4621 (1995).

    Article  ADS  CAS  Google Scholar 

  69. Andrews, S. Fast QC: a quality control tool for high throughput sequence data. http://www.bioinformatics.babraham.ac.uk/projects/fastqc/ (2016).

  70. Callahan, B. J. et al. DADA2: High-resolution sample inference from Illumina amplicon data. Nat. Method 13, 581–583 (2016).

    Article  CAS  Google Scholar 

  71. Lan, Y., Wang, Q., Cole, J. R. & Rosen, G. L. Using the RDP classifier to predict taxonomic novelty and reduce the search space for finding novel organisms. PLoS ONE 7, e32491 (2012).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  72. Quast, C. et al. The SILVA ribosomal RNA gene database project: improved data processing and web-based tools. Nucleic Acids Res. 41, D590–D596 (2013).

    Article  CAS  PubMed  Google Scholar 

  73. McLaren, M. R. Silva SSU taxonomic training data formatted for DADA2. Zenodo https://doi.org/10.5281/zenodo.3986799 (2020).

  74. Wright, E. S. DECIPHER: harnessing local sequence context to improve protein multiple sequence alignment. BMC Bioinformatics 16, 322 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  75. Price, M. N., Dehal, P. S. & Arkin, A. P. FastTree 2—approximately maximum-likelihood trees for large alignments. PLoS ONE 5, e9490 (2010).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  76. R: A Language and Environment for Statistical Computing, v. 4.2.1 (R Foundation for Statistical Computing, 2019).

  77. McMurdie, P. J. & Holmes, S. phyloseq: an R package for reproducible interactive analysis and graphics of microbiome census data. PLoS ONE 8, e61217 (2013).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  78. Oksanen, J. et al. vegan: Community ecology package. https://CRAN.R-project.org/package=vegan (2017).

  79. 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 

  80. Holmes, J. HPCBio/hergenrother-16S-mouse-2022Sept: Initial release. https://doi.org/10.5281/zenodo.10980656 (2024).

  81. 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 

Download references

Acknowledgements

The authors thank L. Li for LC/MS-MS analysis; L. Dirikolu for the pharmacokinetic data analysis; A. Hernandez, C. Wright, M. Band, E. Hogan and J. Drnevich for PacBio sequencing; and R. Rajabi-Toustani and the Core Facilities at the Carl Woese Institute for Genomic Biology for assistance with confocal imaging. We thank the University of Illinois and the NIH (AI136773 and P41-104601) for funding this work. K.A.M. was a member of the NIH Chemistry–Biology Interface Training Grant (T32-GM136629). R.J.U. is supported by an NIH Ruth Kirschstein Award (F31AI161953) and was a NSF predoctoral fellow. This study used computational resources provided by the Delta Advanced Computing and Data Resource, which is supported by the National Science Foundation (award OAC 2005572) and the State of Illinois. Delta is a joint effort of the University of Illinois Urbana-Champaign and its National Center for Supercomputing Applications. This work used the Anvil system at Purdue University through allocation MCA06N060 from the Advanced Cyberinfrastructure Coordination Ecosystem: Services & Support (ACCESS) programme, which is supported by National Science Foundation grants 2138259, 2138286, 2138307, 2137603 and 2138296.

Author information

Authors and Affiliations

Authors

Contributions

K.A.M. and P.J.H. conceived this study. K.A.M. designed and synthesized all novel compounds described and utilized in this Article. K.A.M. and R.J.U. designed and carried out biological experiments. K.A.M. performed aerobic and anaerobic MIC assays, frequency of resistance and time–kill kinetics studies. R.J.U. performed confocal microscopy, co-culture competition experiments, accumulation and frequency of resistance studies. A.K.V., M.S. and P.-C.W. conducted molecular dynamics simulations. J.R.H. and C.J.F. performed bioinformatics and statistical analyses on long read 16S microbiome data. H.Y.L., C.-C.H. and G.W.L. conducted studies with mouse models. K.A.M. and P.J.H. wrote the manuscript with input from R.J.U., A.K.V., M.S., P.-C.W. and E.T. All authors approved the final draft of the manuscript.

Corresponding author

Correspondence to Paul J. Hergenrother.

Ethics declarations

Competing interests

The University of Illinois has filed patents on compounds described in this Article on which K.A.M. and P.J.H. are named inventors.

Peer review

Peer review information

Nature thanks Gerard Wright 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 Previously identified LolCDE inhibitors.

(a) Antibiotic assessment of compounds 1 and 2 against a panel of gram-negative pathogens performed as part of this study. MICs were performed in Mueller Hinton broth per CLSI guidelines and are reported in µg/mL. All experiments were performed in biological triplicate. (b) eNTRy rule parameters (Rotatable bonds, Globularity, Functional Group) of compounds 1 and 2 calculated using eNTRyway33,81. Accumulation determined via previously reported accumulation assay33 and reported in nmol per 1012 colony-forming units (CFUs). Line indicates average accumulation of low accumulating control antibiotics novobiocin, fusidic acid, erythromycin, and rifampicin. Data shown represents the average of three independent experiments with standard deviation of the mean.

Source data

Extended Data Fig. 2 Growth competition of lolamicin resistant mutants generated in E. coli BW25113 as compared to wild-type (WT) E. coli BW25113.

Fitness of E. coli isolates that harbor the mutations in lolC (a) or lolE (b) conferring resistance to lolamicin at concentrations 32-fold above the MIC (64 µg/mL) were evaluated for growth in culture relative to the parental strain. Bacteria were cultured in cation-adjusted Mueller Hinton broth at a starting density of 5 × 103 CFUs/mL and grown for 48 h at 37 °C. At time = 0 hr, 24 hr, and 48 hr, cultures were serially diluted and plated on LB agar and LB agar containing 8 µg/mL lolamicin to quantify number of wild-type and lolamicin resistant mutants. Each competition between mutant and the parental strain was assessed in biological triplicate. Measurements were compared using a two-sample Welch’s t-test (one-tailed test, assuming unequal variance). NS, not significant (P ≥ 0.05); * (P < 0.05); ** (P ≤ 0.01). LolC-E195K t = 24 hr (P = 0.01), t = 48 hr (P = 0.03); LolC-E255D t = 24 hr (P = 0.87), t = 48 hr (P = 0.05); LolC-Q258P t = 24 hr (P = 0.20), t = 48 hr (P = 0.35); LolC-M262I t = 24 hr (P = 0.44), t = 48 hr (P = 0.18); LolC-N265k t = 24 hr (P = 0.79), t = 48 hr (P = 0.55); LolE-D264N t = 24 hr (P = 0.57), t = 48 hr (P = 0.24); LolE-L199P t = 24 hr (P = 0.22), t = 48 hr (P = 0.02); LolE-I206N t = 24 hr (P = 0.16), t = 48 hr (P = 0.42); LolE-F367S t = 24 hr (P = 0.14), t = 48 hr (P = 0.04). ǂLolE F367S displayed morphological changes and smaller colonies compared to wild-type E. coli BW25113.

Source data

Extended Data Fig. 3 Time-kill kinetics of lolamicin against gram-negative pathogens.

(a) The effect of lolamicin and ciprofloxacin on E. coli BW25113 growth. (b) The effect of lolamicin, tetracycline, and ciprofloxacin on K. pneumoniae ATCC 27736 growth. (c) The effect of lolamicin and ciprofloxacin on E. cloacae ATCC 29893 growth. All experiments were performed in biological triplicate and are represented as mean ± s. e. m.

Source data

Extended Data Fig. 4 Lolamicin induces cell swelling in wild-type E. coli and K. pneumoniae but not in lolamicin-resistant mutants.

Confocal microscopy of (a) E. coli BW25113; lolamicin-resistant mutants, (b) BW25113 LolC-N265K, (c) BW25113 LolE-D264N; (d) K. pneumoniae ATCC 27736; and lolamicin-resistant mutants, (e) LolC-Q258L and (f) LolE-V59L. Scale bar is 10 µm. Antibiotics were tested at the following concentrations (3X MIC or just below the solubility limit for lolamicin in resistant mutants): E. coli—DMSO 2%; lolamicin 8 µg/mL for E. coli BW25113, 64 µg/mL for resistant strains; globomycin 24 µg/mL; mecillinam 0.4 µg/mL; aztreonam 0.1 µg/mL. K. pneumoniae—DMSO 2%; lolamicin 3 µg/mL for K. pneumoniae ATCC 27736 or 64 µg/mL for resistant strains; globomycin 64 µg/mL; mecillinam 3 µg/mL; aztreonam 1.5 µg/mL. Cell size (n = 25) in E. coli (g) and K. pneumoniae (h) and resistant mutants was quantified. Length and width were measured in ImageJ and cell area calculated using the area formula for an ellipse (A = π*ab where a = ½ length and b = ½ width). Measurements were compared using two-sample Welch’s t-test (one-tailed test, assuming unequal variance). NS, not significant (P > 0.05); *** (P < 0.0005). E. coli BW25113: lolamicin (P = 3.44 × 10−18), globomycin (P = 1.00 × 10−15); E. coli BW25113 LolC-N265K: lolamicin (P = 0.28), globomycin (P = 4.16 × 10−10); E. coli BW25113 LolE-D264N: lolamicin (P = 0.09), globomycin (P = 4.44 × 10−12). K. pneumoniae ATCC 27736: lolamicin (P = 3.59 × 10−17), globomycin (P = 3.22 × 10−16); K. pneumoniae ATCC 27736 LolC-Q258L: lolamicin (P = 0.22), globomycin (P = 1.26 × 10−8); K. pneumoniae ATCC 27736 LolE-V59L: lolamicin (P = 0.24), globomycin (P = 2.59 × 10−11).

Source data

Extended Data Fig. 5 Conformational landscape of LolCDE.

(a) High-frequency residue/lolamicin contact probability. (b) Probability density of lolamicin’s center of mass locations projected onto two reaction coordinates: distances to BS1 and BS2 (blue). Overlaid color traces show compound 3 unbinding in five simulation replicates, highlighting consistent and immediate unbinding from BS1. (c) Reaction coordinates (RCs) used to project free energy landscape of transporter. Two orientation-based RCs, opening of the TMD periplasmic region (α) and opening of the TMD intracellular region (β), and two distance-based RCs, distance between the nucleotide binding domains (dNBD), and distance between periplasmic domains (dPD), were used for conformation projections. (d) Free energy landscape projected onto RCs. Red dots indicate position of starting Cryo-EM structure (7MDX). (e) Lolamicin-resistant mutations overlap with binding pocket for lipoprotein substrates of LolCDE. Predicted luminal tunnel displayed as gray surface. (f) Molecular rendering of LolC K195-LolE D264 salt bridge interaction. Initial (transparent purple) and final (opaque purple) conformations of LolC loop demonstrate change in binding pocket shape upon mutation. Donor-acceptor heavy atom distance of 2.69 Å is highlighted for final conformation. Density plot of salt-bridge distance between LolC E/K195-LolE D264 from WT and mutant simulations highlights salt bridge formation with E195K peak density. (g) Lolamicin bound in BS1. Conformational sampling of lolamicin rendered as a density with a single conformer (stick). Mutation sensitive LolE residues D264 and I268 (density and stick), and distant Q198 (stick) shown. (h) LolE-N265K mutation relative to lipid bilayer. Also highlighted is LolC-G357 for which the N/K265-door bar distance was measured. Wider distribution for mutant state demonstrates instability of nearby binding pocket. (i) Root-mean squared fluctuation (RMSF) values from replica simulations of LolE residues 357 to 377 calculated from multiple simulation replicas for WT and F367S mutants. F/S367 is marked with a dashed line. Increased RMSF for mutant is apparent.

Extended Data Fig. 6 Bacterial composition of mouse fecal microbiota obtained by full-length 16 s rRNA sequencing at the Class level.

Taxonomic analysis showing bacterial population shifts over a 31-day period before (Day 0) and after (Day 7, Day 10, and Day 31) administration of antibiotic. CD-1 mice were treated with vehicle (20% DMSO, 30% water, 50% PEG400, n = 6 biologically independent animals) or compound (clindamycin, 100 mg/kg; amoxicillin, 100 mg/kg; or lolamicin, 200 mg/kg; n = 6 biologically independent animals for each compound) twice a day for three days via oral gavage.

Extended Data Fig. 7 Bacterial composition of mouse fecal microbiota obtained by full-length 16 s rRNA sequencing at the Order level.

Taxonomic analysis showing bacterial population shifts over a 31-day period before (Day 0) and after (Day 7, Day 10, and Day 31) administration of antibiotic. CD-1 mice were treated with vehicle (20% DMSO, 30% water, 50% PEG400, n = 6 biologically independent animals) or compound (clindamycin, 100 mg/kg; amoxicillin, 100 mg/kg; or lolamicin, 200 mg/kg; n = 6 biologically independent animals for each compound) twice a day for three days via oral gavage.

Extended Data Fig. 8 PCoA ordination of Bray-Curtis dissimilarity values for samples before (day 0) and after (day 7, day 10) lolamicin, amoxicillin, and clindamycin administration.

Vehicle-treated samples are included as negative controls. Samples are grouped by day and color coded as shown in the key. Points represent individual samples for each grouping. CD-1 mice were treated with vehicle (20% DMSO, 30% water, 50% PEG400, n = 6 biologically independent animals) or compound (clindamycin, 100 mg/kg; amoxicillin, 100 mg/kg; or lolamicin, 200 mg/kg; n = 6 biologically independent animals for each compound) twice a day for three days via oral gavage.

Supplementary information

Supplementary Information

This file contains synthetic methods and NMR spectra for reported compounds and Supplementary Tables 1–12.

Reporting Summary

Supplementary Data (source data Supplementary Table 4).

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

Muñoz, K.A., Ulrich, R.J., Vasan, A.K. et al. A Gram-negative-selective antibiotic that spares the gut microbiome. Nature 630, 429–436 (2024). https://doi.org/10.1038/s41586-024-07502-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-024-07502-0

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