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:

Minimizing higher-order aggregation maximizes iron mobilization by small molecules

Abstract

The natural product hinokitiol mobilizes iron across lipid bilayers at low concentrations and restores hemoglobinization in iron transporter protein-deficient systems. But hinokitiol fails to similarly mobilize iron at higher concentrations, limiting its uses in chemical biology and medicine. Here we show that at higher concentrations, hinokitiol3:Fe(III) complexes form large, higher-order aggregates, leading to loss of transmembrane iron mobilization. Guided by this understanding and systematic structure–function studies enabled by modular synthesis, we identified FeM-1269, which minimally aggregates and dose-dependently mobilizes iron across lipid bilayers even at very high concentrations. In contrast to hinokitiol, FeM-1269 is also well-tolerated in animals at high doses for extended periods of time. In a mouse model of anemia of inflammation, FeM-1269 increases serum iron, transferrin saturation, hemoglobin and hematocrit. This rationally developed iron-mobilizing small molecule has enhanced potential as a molecular prosthetic for understanding and potentially treating iron transporter deficiencies.

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: Constitutional isomers with divergent mobilization and higher-order aggregation profiles.
Fig. 2: Structure–activity study of alkyl tropolones to identify favorable liposome and Caco-2 iron mobilization profiles.
Fig. 3: Further structure–activity study to identify and validate FeM-1269 as an optimized iron mobilizer.
Fig. 4: FeM-1269 is better tolerated than hinokitiol or GT in rats.
Fig. 5: High exposures of FeM-1269 do not disturb trace metal homeostasis.
Fig. 6: FeM-1269 promotes restored iron mobilization in AoI.

Similar content being viewed by others

Data availability

All data are available in the main text, extended data or supplementary materials. Source data are available through the publisher for all data in the main text and extended data figures. Any other data are available from the author(s) upon reasonable request. Crystallographic information has been deposited at the Cambridge Crystallographic Data Center 2088858 for the GT3:Fe(III) structure disclosed here. Source data are provided with this paper.

References

  1. Weiss, G., Ganz, T. & Goodnough, L. T. Anemia of inflammation. Blood 133, 40–50 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Ganz, T. Anemia of inflammation. N. Engl. J. Med. 381, 1148–1157 (2019).

    Article  CAS  PubMed  Google Scholar 

  3. Rivera, S. & Ganz, T. Animal models of anemia of inflammation. Semin. Hematol. 46, 351–357 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Nemeth, E. et al. IL-6 mediates hypoferremia of inflammation by inducing the synthesis of the iron regulatory hormone hepcidin. J. Clin. Investig. 113, 1271–1276 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Nemeth, E. et al. Hepcidin regulates cellular iron efflux by binding to ferroportin and inducing its internalization. Science 306, 2090–2093 (2004).

    Article  CAS  PubMed  Google Scholar 

  6. Wrighting, D. M. & Andrews, N. C. Interleukin-6 induces hepcidin expression through STAT3. Blood 108, 3204–3209 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Nicolas, G. et al. The gene encoding the iron regulatory peptide is regulated by anemia, hypoxia, and inflammation. J. Clin. Investig. 110, 1037–1044 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Kim, A. et al. A mouse model of anemia of inflammation: complex pathogenesis with partial dependence on hepcidin. Blood 123, 1129–1136 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Roy, C. N. et al. Hepcidin antimicrobial peptide transgenic mice exhibit features of the anemia of inflammation. Blood 109, 4038–4044 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Langdon, J. M. et al. Hepcidin-dependent and hepcidin-independent regulation of erythropoiesis in a mouse model of anemia of chronic inflammation. Am. J. Hematol. 89, 470–479 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Cazzola, M. et al. Defective iron supply for erythropoiesis and adequate endogenous erythropoietin production in the anemia associated with systemic-onset juvenile chronic arthritis. Blood 87, 4824–4830 (1996).

    Article  CAS  PubMed  Google Scholar 

  12. Eschbach, J. W. et al. Recombinant human erythropoietin in anemic patients with end-stage renal disease. Ann. Intern. Med. 111, 992–1000 (1989).

    Article  CAS  PubMed  Google Scholar 

  13. Egrie, J. C. & Browne, J. K. Development and characterization of a novel erythropoiesis stimulating protein (NESP). Br. J. Cancer 84, 3–10 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Provenzano, R. et al. Ferumoxytol as an intravenous iron replacement therapy in hemodialysis patients. Clin. J. Am. Soc. Nephrol. 4, 386–393 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Gupta, N. & Wish, J. B. Hypoxia-inducible factor prolyl hydroxylase inhibitors: a potential new treatment for anemia in patients with CKD. Am. J. Kidney Dis. 69, 815–826 (2017).

    Article  CAS  PubMed  Google Scholar 

  16. Food and Drug Administration. FDA Roxadustat briefing document: Roxadustat; NDA 213805. Cardiovascular and Renal Drugs Advisory Committee Meeting (2021). www.fda.gov/media/150728/download

  17. Rostoker, G. & Vaziri, N. D. Risk of iron overload with chronic indiscriminate use of intravenous iron products in ESRD and IBD populations. Heliyon 5, e02045 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  18. Sasu, B. J. Antihepcidin antibody treatment modulates iron metabolism and is effective in a mouse model of inflammation-induced anemia. Blood 115, 3616–3624 (2010).

    Article  CAS  PubMed  Google Scholar 

  19. Grillo, A. S. et al. Restored iron transport by a small molecule promotes absorption and hemoglobinization in animals. Science 356, 608–616 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Yien, Y. Y. et al. FAM210B is an erythropoietin target and regulates erythroid heme synthesis by controlling mitochondrial iron import and ferrochelatase activity. J. Biol. Chem. 293, 19797–19811 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Rondelli, C. M. et al. The ubiquitious mitochondrial protein unfoldase CLPX regulates erythroid heme synthesis by control of iron utilization and heme synthesis enzyme activation and turnover. J. Biol. Chem. 297, 100972 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Rider, R. C. & Hoffbrand, A. V. The role of deferiprone in iron chelation. N. Engl. J. Med. 379, 2140–2150 (2018).

    Article  Google Scholar 

  23. Propper, R. D. et al. Continuous subcutaneous administration of deferoxamine in patients with iron overload. N. Engl. J. Med. 297, 418–423 (1977).

    Article  CAS  PubMed  Google Scholar 

  24. Kontoghiorghes, G. J. et al. Effective chelation of iron in β thalassemia with the oral chelator 1,2-dimethyl-3-hydroxypyrid-4-one. Br. J. Med (Clin. Res. Ed.) 295, 1509–1512 (1987).

    Article  CAS  Google Scholar 

  25. Ihnat, P. M., Vennerstrom, J. L. & Robinson, D. H. Solution equilibria of deferoxamine amides. J. Pharm. Sci. 91, 1733–1741 (2002).

    Article  CAS  PubMed  Google Scholar 

  26. Ekaputri, S. E. et al. A small molecule redistributes iron in ferroportin deficient mice and patient-derived primary macrophages. Proc. Natl Acad. Sci. USA 119, e2121400119 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Morita, Y. et al. Biological activity of α-thujaplicin, the minor component of Thujopsis dolabrata SIEB. et ZUCC. var. hondai MAKINO. Biol. Pharm. Bull. 24, 607–611 (2001).

    Article  CAS  PubMed  Google Scholar 

  28. Morita, Y. et al. Biological activity of α-dolabrin, β-thujaplicin, and 4-acetyltropolone, hinokitiol-related compounds. Biol. Pharm. Bull. 27, 1666–1669 (2004).

    Article  CAS  PubMed  Google Scholar 

  29. Li, J. et al. Synthesis of many different types of organic molecules using one automated process. Science 347, 1221–1226 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Lehmann, J. W. et al. Axial shielding of Pd(II) complexes enables perfect stereoretention in Suzuki–Miyaura cross-coupling of Csp3 boronic acids. Nat. Commun. 10, 1263 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  31. Blair, D. J. et al. Automated iterative Csp3–C bond formation. Nature 604, 92–97 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Coan, K. E. D. & Shoichet, B. K. Stoichiometry and physical chemistry of promiscuous aggregation-based inhibitors. J. Am. Chem. Soc. 130, 9606–9612 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Stetefelt, J., McKenna, S. A. & Patel, T. R. Dynamic light scattering: a practical guide and applications in biomedical sciences. Biophys. Rev. 8, 409–427 (2016).

    Article  Google Scholar 

  34. Anderson, T. M. et al. Amphotericin forms an extramembranous and fungicidal sterol sponge. Nat. Chem. Biol. 10, 400–406 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Fedoryński, M. Syntheses of gem-dihalocyclopropanes and their use in organic synthesis. Chem. Rev. 103, 1099–1132 (2003).

    Article  PubMed  Google Scholar 

  36. Banwell, M. G. & Onrust, R. A versatile strategy for the synthesis of tropolones. Tetrahedron Lett. 26, 4543–4546 (1985).

    Article  CAS  Google Scholar 

  37. Nicolaou, K. C., Yu, R., Lu, Z. & Alvarez, F. G. Total synthesis of gukulenin B via sequential tropolone functionalizations. J. Am. Chem. Soc. 144, 5190–5196 (2022).

    Article  CAS  PubMed  Google Scholar 

  38. Imao, D., Glasspoole, B. W., Laberge, V. S. & Crudden, C. M. Cross coupling reactions of chiral secondary organoboronic esters with retention of configuration. J. Am. Chem. Soc. 131, 5024–5025 (2009).

    Article  CAS  PubMed  Google Scholar 

  39. Cherney, A. H., Kadunce, N. T. & Reisman, S. E. Enantioselective and enantiospecific transition-metal-catalyzed cross-coupling reactions of organometallic reagents to construct C–C bonds. Chem. Rev. 115, 9587–9652 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Morstein, J. et al. Medium-length lipid conjugation facilitates cell-permeability and bioactivity. J. Am. Chem. Soc. 144, 18532–18544 (2022).

    Article  CAS  PubMed  Google Scholar 

  41. Daina, A., Michielin, O. & Zoete, V. SwissADME: a free web tool to evaluate pharmacokinetics, drug-likeness and medicinal chemistry friendliness of small molecules. Sci. Rep. 7, 42717 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  42. Hidalgo, I. J., Raub, T. J. & Borchardt, R. T. Characterization of the human colon carcinoma cell line (Caco-2) as a model system for intestinal epithelial permeability. Gastroenterology 96, 736–749 (1989).

    Article  CAS  PubMed  Google Scholar 

  43. Lee, Y.-S. et al. Hinokitiol inhibits cell growth through induction of S-phase arrest and apoptosis in human colon cancer cells and suppresses tumor growth in a mouse xenograft experiment. J. Nat. Prod. 27, 2195–2202 (2013).

    Article  Google Scholar 

  44. Cho, Y. M. et al. A 13-week subchronic toxicity study of hinokitiol administered in the diet to F344 rats. Food Chem. Toxicol. 49, 1782–1786 (2011).

    Article  CAS  PubMed  Google Scholar 

  45. Imai, N. et al. Lack of hinokitiol (β-thujaplicin) carcinogenicity in F344/DuCrj rats. J. Toxicol. Sci. 31, 357–370 (2006).

    Article  CAS  PubMed  Google Scholar 

  46. Ema, M., Harazono, A., Fujii, S. & Kawashima, K. Evaluation of developmental toxicity of β-thujaplicin (hinokitiol) following oral administration during organogenesis in rats. Food Chem. Toxicol. 42, 465–470 (2004).

    Article  CAS  PubMed  Google Scholar 

  47. Kunes, M., Kvetina, J. & Bures, J. Type and distribution of indomethacin-induced lesions in the gastrointestinal tract in rat. Neuro Endocrinol. Lett. 30, 96–100 (2009).

    CAS  PubMed  Google Scholar 

  48. Rachmilewitz, D. et al. Inflammatory mediators of experimental colitis in rats. Gastroenterology 97, 326–337 (1989).

    Article  CAS  PubMed  Google Scholar 

  49. Paulson, R. F., Hariharan, S. & Little, J. Stress erythropoiesis: definitions and models for its study. Exp. Hematol. 89, 43–54 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Brissot, P. et al. Haemochromatosis. Nat. Rev. Dis. Primers 4, 18016 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  51. Zohn, I. E. et al. The flatiron mutation in mouse ferroportin acts as a dominant negative to cause ferroportin disease. Blood 109, 4174–4180 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Garbers, C., Heink, S., Korn, T. & Rose-John, S. Interleukin-6: designing specific therapeutics for a complex cytokine. Nat. Rev. Drug Discov. 17, 395–412 (2018).

    Article  CAS  PubMed  Google Scholar 

  53. Guralnik, J. M., Eisenstaedt, R. S., Ferrucci, L., Klein, H. G. & Woodman, R. C. Prevalence of anemia in persons 65 years and older in the United States: evidence for a high rate of unexplained anemia. Blood 15, 2263–2268 (2004).

    Article  Google Scholar 

  54. Deng, Q. et al. Lower hemoglobin levels in patients with Parkinson’s disease are associated with disease severity and iron metabolism. Brain Res. 15, 145–151 (2017).

    Article  Google Scholar 

  55. Roy, N. C. & Andrews, N. C. Recent advances in disorders of iron metabolism: mutations, mechanisms and modifiers. Hum. Mol. Genet. 10, 2181–2186 (2001).

    Article  CAS  PubMed  Google Scholar 

  56. Jędrzejewska, H. & Szumna, A. Making a left or right choice: chiral self-sorting as a tool for the formation of discrete complex structures. Chem. Rev. 117, 4863–4899 (2017).

    Article  PubMed  Google Scholar 

  57. Kong, W.-N. et al. Sex differences in iron status and hepcidin expression in rats. Biol. Trace Elem. Res. 160, 258–267 (2014).

    Article  CAS  PubMed  Google Scholar 

  58. Muraglia, K. A. et al. Small molecule ion channel restores host defenses in cystic fibrosis airway epithelia. Nature 567, 405–408 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Chorghade, R. S. et al. Amphotericin B induces epithelial voltage responses in people with cystic fibrosis. J. Cyst. Fibros. 20, 540–550 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  60. Guthrie, L. M. et al. Elesclomol alleviates menkes pathology and mortality by escorting Cu to cuproenzymes in mice. Science 368, 620–625 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Suzuki, K. et al. A synthetic synaptic organizer protein restores glutamatergic neuronal circuits. Science 369, eabb4853 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Su, T. A. et al. A modular ionophore platform for liver-directed copper supplementation in cells and animals. J. Am. Chem. Soc. 140, 13764–13774 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Chao, J. et al. Tropolone derivatives and tautomers thereof for iron regulation in animals. patentscope.wipo.int/search/en/detail.jsf?docId=WO2021076945 (2021).

  64. Chao, J. et al. Tropolone derivatives and tautomers thereof for iron regulation in animals. patentscope.wipo.int/search/en/detail.jsf?docId=WO2021076938&_cid=P21-LU0T05-53662-1 (2021).

  65. Stookey, L. L. Ferrozine—a new spectrophotometric reagent for iron. Anal. Chem. 42, 779–781 (1970).

    Article  CAS  Google Scholar 

  66. Tandy, S. et al. Nramp2 expression is associated with pH-dependent iron uptake across the apical membrane of human intestinal Caco-2 cells. J. Biol. Chem. 275, 1023–1029 (2000).

    Article  CAS  PubMed  Google Scholar 

  67. Lesjak, M. et al. Quercetin inhibits intestinal iron absorption and ferroportin transporter expression in vivo and in vitro. PLoS ONE 9, e102900 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  68. Seo, Y. A. & Wessling-Resnick, M. Ferroportin deficiency impairs manganese metabolism in flatiron mice. FASEB J. 29, 2726–2733 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Seo, Y. A., Elkhader, J. A. & Wessling-Resnick, M. Distribution of manganese and other biometals in flatiron mice. Biometals 29, 147–155 (2016).

    Article  CAS  PubMed  Google Scholar 

  70. Sanchez, A. et al. Stress erythropoiesis in atherogenic mice. Sci. Rep. 10, 18469 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank R.S. Chorghade for helpful discussions around liposome experiments, S. Robinson (Beckman Imaging Technology Group) for assistance with TEM imaging, T. Woods (George R. Clark X-Ray Facility and 3M Materials Laboratory) for X-ray structure determination, J. Katzenellenbogen for the use of a liquid scintillation counter, A.S. Grillo, A.G. Cioffi and D.J. Blair for their work on the early synthesis of tropolone derivatives and C. Murphy for the use of a Zetasizer Nano for DLS studies. TEM images were taken at the Beckman Institute for Advanced Science & Technology Imaging Technology Group (ITG) Microscopy Suite on a Thermo Fisher Scientific FEI Tecnai G2 F20 S-TWIN 200 kV STEM. The Bruker 500 MHz nuclear magnetic resonance (NMR) spectrometer was obtained with the financial support of the Roy J. Carver Charitable Trust, Muscatine, Iowa, USA. Metal analysis was performed at the Northwestern University Quantitative Bio-element Imaging Center and was funded by the Chicago Biomedical Consortium with support from the Searle Funds at The Chicago Community Trust and supported by the Office Of The Director, National Institutes of Health of the National Institutes of Health under award S10OD020118. Portions of this work were supported by the National Institutes of Health (T32GM070421 (to A.D.B.), R01HL140526 (to Y.A.S. and M.D.B.) and R35GM118185 (to M.D.B.)).

Author information

Authors and Affiliations

Authors

Contributions

A.D.B., A.M.S.M. and M.D.B. designed the study. A.D.B. performed liposome efflux, Caco-2, DLS, TEM, X-ray crystallographic analysis and statistical analyses. A.M.S.M. and K.J.G. performed Caco-2 studies. A.D.B. and P.-J.C. synthesized and evaluated the initial set of alkyl tropolonoids. J.C. and D.M.M. conducted the medicinal chemistry campaign that identified FeM-1269. N.M.S., T.G. and A.O. designed, coordinated and performed animal studies and performed statistical analyses. C.K.R. and S.T.B. performed mechanistic DLS studies. C.K.R. and S.E. performed studies comparing iron chelators and iron mobilizers. E.-K.C., L.A. and Y.A.S. performed studies in the flatiron mouse anemia model. T.F. performed the mouse study assessing stress erythropoiesis, assisted by S.E. and K.J. G. S.J.H., Y.A.S. and M.D.B. oversaw the work described in this study. A.D.B. and M.D.B. wrote the paper, which was reviewed by all authors.

Corresponding authors

Correspondence to Young-Ah Seo or Martin D. Burke.

Ethics declarations

Competing interests

M.D.B. is an inventor on the US patent application pub. 20180263926, ‘Restoring Physiology in Iron-Deficient Organisms Using Small Molecules’, which is directed to the use of small-molecule iron transporters to treat deficiencies of iron-transport proteins. A.D.B., A.M.S.M. and M.D.B. are inventors on the US patent application pub. 20210163393, ‘Hinokitiol analogs, methods of preparing and pharmaceutical compositions thereof’, which is directed to analogs of hinokitiol, preparation and formulation. These patent applications, their foreign counterparts and others have been licensed by Cajal Neuroscience, for which M.D.B. is a scientific advisor and shareholder. Ambys Medicines has filed patent applications directed to tropolone derivatives and tautomers thereof for iron regulation in animals. J.C., N.M.S., D.M.M., T.G., A.O. and S.J.H. were full-time employees of Ambys Medicines when these data were generated. These patents are currently owned by Cajal Neuroscience. The other authors declare no competing interests.

Peer review

Peer review information

Nature Chemical Biology thanks Christopher Chang and the other, anonymous, reviewers 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 In vitro iron mobilization and aggregation by tropolonoids.

a, Time course of iron efflux from liposomes with different concentrations of hinokitiol, GT or FeM-1269 (n = 3; representative experiment shown). b, Rates of iron efflux from liposomes at each concentration of compounds, calculated from data in a–c. c, DLS derived count rate measured in different concentrations of MES/TRIS buffer, pH 7.0 (n = 3; representative experiment shown). d, DLS derived count rate measured at different pH of 50 mM MES/TRIS buffer (n = 3; representative experiment shown). e, DLS derived count rate measured with different buffering agents at 50 mM of buffer (n = 3; representative experiment shown). f, DLS derived count rate of hinokitiol2:Cu(II), GT2:Cu(II) and FeM-12692:Cu(II) (n = 3; representative experiment shown). g, Additional representative images from 90 μM of hino3:Fe(III), GT3:Fe(III) or FeM-12693:Fe(III) in buffer (n = 50, 45, 51 images for hino, GT and FeM-1269, respectively; images were taken from three independent sample preparations; scale bar is 500 nm). h, Iron efflux by tropolone derivatives from shDMT1 Caco-2 epithelia at 1 μM (n = 3, representative experiment shown). i, Iron efflux by tropolone derivatives from shDMT1 Caco-2 epithelia at 5 μM (n = 3, representative experiment shown).

Source data

Extended Data Fig. 2 Physicochemical parameters of tropolone iron mobilizer library.

a, Physicochemical parameters of tropolones relative to performance in liposomes, defined by Elip. b, Physicochemical parameters of tropolones relative to iron mobilized in shDMT1-Caco-2 differentiated gut epithelia after 4 hours at 3 μM, tracked by 55FeCl3 (n = 3; experiments were performed at least twice). c, Physicochemical parameters of tropolones relative to iron mobilized in shDMT1-Caco-2 differentiated gut epithelia after 4 hours at 10 μM, tracked by 55FeCl3 (n = 3; experiments were performed at least twice). All parameters were computed in SwissADME.

Source data

Extended Data Fig. 3 Physicochemical parameters and chemical structures of highly effective iron mobilizers.

a–f, Physicochemical parameters of top tropolones in each assay for iron mobilization (n = 20 tropolones). Extracted from the datasets displayed in Extended Data Fig. 2. g, Chemical structures of the top 20 tropolone iron mobilizers in shDMT1-Caco-2 epithelia at 10 μΜ are largely substituted with short alkyl- or alkoxy-side chain substituents. Of note, (+)-6 and (−)-6 are isolated during the preparation of (±)-6 and were independently evaluated. See the Supplementary Note for details on chiral separation. Error bars indicate SEM.

Source data

Extended Data Fig. 4 FeM-1269 has a favorable drug-like profile.

a, Efflux of Fe(III) after 30 minutes, tracked by incorporation into ferrozine (n = 3; representative experiment shown). b, Iron efflux from shDMT1 Caco-2 at several concentrations (n = 3; representative experiment shown). c, Fe(III) efflux from FPN1-KO RAW264.7 macrophages after 2 hours, tracked by 55FeCl3 (n = 3; representative experiment shown). Error bars indicate SEM. d, Dashboard of in vitro ADME studies on FeM-1269. e, Stoichiometric competition experiment with 1 mM of each divalent metal and 1 mM of the indicated tropolone. Tropolone-bound metal was measured by ICP-MS after equilibration and extraction with ethyl acetate (n = 3 independent samples). f, Serum plasma content in mice of tropolonoids after single-dose of 10 mg/kg in water for injection (n = 3 animals per group). g, Calculated pharmacokinetic parameters of data in c. h, Maximum tolerated dose finding and pharmacokinetic study in mice of orally administered FeM-1269 (n = 2 animals per group) – MTD = 1000 mg/kg. i, Correlation between exposure and administered dose, using data in Fig. 4d and the study in f. j, Calculated pharmacokinetic calculations for MTD study in e (n = 2 per group). k, Results from micronucleus testing at MTD of FeM-1269 of 1000 mg/kg/day – (n = 5, 5, 4 animals for vehicle treated, 24 hours post-FeM-1269 and 48 hours post-FeM-1269, respectively). l, Pharmacokinetic study of FeM-1269 in rats (n = 3 animals per group). m, Pharmacokinetic study of FeM-1269 in beagle dogs (n = 3 animals per group). Detailed pharmacokinetic calculations are available in Supplementary Table 3 for the studies in l,m. Error bars indicate SEM.

Source data

Extended Data Fig. 5 Additional outcomes from two-week toxicology studies with hinokitiol and GT.

a, Animal weights from two-week daily administration of hinokitiol at 10, 30 and 100 mg/kg/day (n = 6 animals per group). b, Pharmacokinetic study to establish bioavailability and approximate exposure to hinokitiol (n = 3 animals per group). c, Pharmacokinetics of hinokitiol administered on day 1 to rats (n = 3 animals per group). d, Pharmacokinetics of hinokitiol administered on day 15 to rats (n = 3 animals per group). e, Pharmacokinetic calculations from studies in c,d. f, Representative H&E staining of rat myocardium from an animal treated with 30 mg/kg of hinokitiol. g, Representative H&E staining of rat myocardium from an animal treated with 100 mg/kg of hinokitiol. h, Representative H&E staining of rat myocardium from an animal treated with 100 mg/kg of hinokitiol. i, Rat weights from two-week daily administration of GT at 15, 45 and 105 mg/kg/day (n = 6 animals per group). j, Pharmacokinetic study to establish bioavailability and approximate exposure to GT (n = 3 animals per group). k, Pharmacokinetic study of serum concentrations of GT on day 1 of the study (n = 3 animals per group). l, Pharmacokinetic study of serum concentrations of GT on day 15 of the study (n = 3 animals per group). m, Pharmacokinetic calculations from studies with GT in c,d. n, Representative H&E staining of rat myocardium from an animal treated with 45 mg/kg of GT. o, Representative H&E staining of rat myocardium from an animal treated with 105 mg/kg of GT. p, Representative DAB-Perl’s staining of rat myocardium from an animal treated with 105 mg/kg of GT. Arrows indicate features of interest. For panels f–h, n–p, the scale bars indicate 200 μm, and a minimum of three animals were processed and imaged for each treatment group. Pharmacokinetic calculations were performed in WinNonlin. Error bars indicate SEM.

Source data

Extended Data Fig. 6 Additional data from two-week rat toxicology study with FeM-1269.

a, Animal weights from two-week daily administration of FeM-1269 at 30, 100 and 300 mg/kg/day in male rats (n = 3 animals per group). b, Animal weights from two-week daily administration of FeM-1269 at 30, 100 and 300 mg/kg/day in female rats (n = 3 animals per group). c, Pharmacokinetic study of serum concentrations of FeM-1269 on day 1 of the study in male rats (n = 2 animals per group). d, Pharmacokinetic study of serum concentrations of FeM-1269 on day 15 of the study in males (n = 2 animals per group). e, Pharmacokinetic study of serum concentrations of FeM-1269 on day 1 of the study in females (n = 2 animals per group). f, Pharmacokinetic study of serum concentrations of FeM-1269 on day 15 of the study in females (n = 2 animals per group). g, Calculated pharmacokinetic parameters from studies in c–f. Error bars indicate SEM.

Source data

Extended Data Fig. 7 Additional data from three-month mouse toxicology studies with FeM-1269 and stress erythropoiesis follow-up.

a, Male weights with daily administration of FeM-1269 (n = 6 animals per group). b, Female weights with daily administration of FeM-1269 (n = 6 animals per group). c,d, Histopathological analysis of tissues following FeM-1269 administration for 90 days in male and female mice (n = 6 animals per group). Only the tissues in which findings were observed were included in these graphs. Observation of the following tissues yielded no histopathological findings: brain, skeletal muscle, small intestine, spleen or thymus. Statistics in c,d were performed using two-way ANOVA with Dunnett’s multiple comparisons test. e, Female mouse weights following treatment with vehicle (0.5% HPMC) or 150 mg/kg/day of FeM-1269 (n = 12 animals per group). f, Spleen iron as assessed by ICP-MS (n = 12 per group, p = 0.0437). g, Histopathology analysis of mouse spleens using Prussian blue staining (n = 12 per group, p = 0.1114). h, Counts of bone marrow burst-forming unit, erythrocytes (BFU) (n = 12 per group, p = 0.4848). i, Counts of bone marrow colony-forming unit, erythrocytes (CFU) (n = 12 per group, p = 0.6243). j, Counts of spleen-derived BFU (n = 12 per group, p = 0.9999). k, Counts of bone marrow burst-forming unit, erythrocytes (CFU) (n = 12 per group, p = 0.0934). Statistics in f–k were performed using two-sided unpaired t-tests with Welch’s correction. For p-values for c,d, please see the source data file. Error bars represent SEM.

Source data

Extended Data Fig. 8 Studies of FeM-1269 administration in flatiron mice.

In a–d, animals were administered vehicle or increasing doses of FeM-1269 via oral gavage and sacrificed 3 hours later for data collection (n = 6 animals for all groups). a, Duodenum non-heme iron levels. b, Spleen non-heme iron levels. c, Liver non-heme iron levels. d, Transferrin saturation. In e–k, animals were given 10 mg/kg of either hinokitiol, FeM-1269 or deferiprone daily for seven days, then sacrificed three hours after the final dose for data collection (n = 6 animals for WT + vehicle, FeM-1269 10 mg/kg; n = 5 for other groups). e, Duodenum non-heme iron levels (n = 6 animals for WT + vehicle, FeM-1269 10 mg/kg; n = 5 for other groups). f, Spleen non-heme iron levels. g, Liver non-heme iron levels. h, Transferrin saturation levels. i, Total serum iron levels. j, Hemoglobin. k, Hematocrit. Statistics were performed using one-way ANOVA with Tukey’s multiple comparisons test between each group and the vehicle treated group. For p-values, please see the source data file. Error bars indicate SEM. Where displayed, *p < 0.05 relative to vehicle controls.

Source data

Extended Data Fig. 9 Additional data from anemia of inflammation study in Fig. 6.

In a–f, mice were treated as described in the supporting information section ‘Anemia of Inflammation Dose-Escalation Study’ (n = 10 animals for Sham and TO + bleed groups; n = 12 for all other groups). a, Weight of mice employed in the study. b, Calculated daily exposure to FeM-1269 based on pharmacokinetic data in Extended Data Fig. 4i. c, Transferrin saturation following daily administrations of hinokitiol (predicted area under the curve (AUC) 9560 ng/mL * hr). d, Total serum iron following daily administration of hinokitiol. e, Hemoglobin following daily administration of hinokitiol. f, Hematocrit following daily administration of hinokitiol. Sham and vehicle data in c–f are replicated from Fig. 6. Statistics were analyzed in Prism using one-way ANOVA with Dunnett’s or Tukey’s multiple comparisons tests as appropriate. For p-values, please see the source data file. Error bars indicate SEM. Where displayed, *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001 relative to vehicle controls or as otherwise specified.

Source data

Supplementary information

Supplementary Information

Supplementary Figs. 1–25, Supplementary Tables 1–51 and Supplementary Note.

Reporting Summary

Supplementary Data 1

Supporting data for Supplementary Fig. 7.

Supplementary Data 2

Supporting data for Supplementary Fig. 8.

Supplementary Data 3

Supporting data for Supplementary Fig. 9.

Supplementary Data 4

Supporting data for Supplementary Fig 10.

Supplementary Data 5

Supporting data for Supplementary Fig. 11.

Supplementary Data 6

Supporting data for Supplementary Fig. 12.

Supplementary Data 7

Supporting data for Supplementary Fig 13.

Supplementary Data 8

Supporting data for Supplementary Fig 14.

Supplementary Data 9

Supporting data for Supplementary Fig. 15.

Supplementary Data 10

Supporting data for Supplementary Fig 16.

Supplementary Data 11

Supporting data for Supplementary Fig. 17.

Supplementary Data 12

Supporting data for Supplementary Fig. 18.

Supplementary Data 13

Supporting data for Supplementary Fig 19.

Supplementary Data 14

Supporting data for Supplementary Fig. 20.

Supplementary Data 15

Supporting data for Supplementary Fig. 21.

Supplementary Data 16

Supporting data for Supplementary Fig. 22.

Supplementary Data 17

Supporting data for Supplementary Fig. 23.

Supplementary Data 18

Supporting data for Supplementary Fig. 24.

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 Fig. 6

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.

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

Blake, A.D., Chao, J., SantaMaria, A.M. et al. Minimizing higher-order aggregation maximizes iron mobilization by small molecules. Nat Chem Biol (2024). https://doi.org/10.1038/s41589-024-01596-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1038/s41589-024-01596-3

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