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.

  • Letter
  • Published:

Itaconate is an anti-inflammatory metabolite that activates Nrf2 via alkylation of KEAP1

Abstract

The endogenous metabolite itaconate has recently emerged as a regulator of macrophage function, but its precise mechanism of action remains poorly understood1,2,3. Here we show that itaconate is required for the activation of the anti-inflammatory transcription factor Nrf2 (also known as NFE2L2) by lipopolysaccharide in mouse and human macrophages. We find that itaconate directly modifies proteins via alkylation of cysteine residues. Itaconate alkylates cysteine residues 151, 257, 288, 273 and 297 on the protein KEAP1, enabling Nrf2 to increase the expression of downstream genes with anti-oxidant and anti-inflammatory capacities. The activation of Nrf2 is required for the anti-inflammatory action of itaconate. We describe the use of a new cell-permeable itaconate derivative, 4-octyl itaconate, which is protective against lipopolysaccharide-induced lethality in vivo and decreases cytokine production. We show that type I interferons boost the expression of Irg1 (also known as Acod1) and itaconate production. Furthermore, we find that itaconate production limits the type I interferon response, indicating a negative feedback loop that involves interferons and itaconate. Our findings demonstrate that itaconate is a crucial anti-inflammatory metabolite that acts via Nrf2 to limit inflammation and modulate type I interferons.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Itaconate activates Nrf2.
Figure 2: Itaconate alkylates cysteines.
Figure 3: OI limits IL-1β in an Nrf2-dependent manner and protects against LPS lethality.
Figure 4: A feedback loop exists between itaconate and IFN-β.

Similar content being viewed by others

References

  1. Michelucci, A. et al. Immune-responsive gene 1 protein links metabolism to immunity by catalyzing itaconic acid production. Proc. Natl Acad. Sci. USA 110, 7820–7825 (2013)

    Article  CAS  ADS  Google Scholar 

  2. Strelko, C. L. et al. Itaconic acid is a mammalian metabolite induced during macrophage activation. J. Am. Chem. Soc. 133, 16386–16389 (2011)

    Article  CAS  Google Scholar 

  3. Lampropoulou, V. et al. Itaconate links inhibition of succinate dehydrogenase with macrophage metabolic remodeling and regulation of inflammation. Cell Metab. 24, 158–166 (2016)

    Article  CAS  Google Scholar 

  4. Mills, E. L. et al. Succinate dehydrogenase supports metabolic repurposing of mitochondria to drive inflammatory macrophages. Cell 167, 457–470 (2016)

    Article  CAS  Google Scholar 

  5. Hayes, J. D. & Dinkova-Kostova, A. T. The Nrf2 regulatory network provides an interface between redox and intermediary metabolism. Trends Biochem. Sci. 39, 199–218 (2014)

    Article  CAS  Google Scholar 

  6. Brennan, M. S. et al. Dimethyl fumarate and monoethyl fumarate exhibit differential effects on KEAP1, NRF2 activation, and glutathione depletion in vitro. PLoS One 10, e0120254 (2015)

    Article  Google Scholar 

  7. ElAzzouny, M. et al. Dimethyl itaconate is not metabolized into itaconate intracellularly. J. Biol. Chem. 292, 4766–4769 (2017)

    Article  CAS  Google Scholar 

  8. Kobayashi, E. H. et al. Nrf2 suppresses macrophage inflammatory response by blocking proinflammatory cytokine transcription. Nat. Commun. 7, 11624 (2016)

    Article  CAS  ADS  Google Scholar 

  9. Lee, J. M., Calkins, M. J., Chan, K., Kan, Y. W. & Johnson, J. A. Identification of the NF-E2-related factor-2-dependent genes conferring protection against oxidative stress in primary cortical astrocytes using oligonucleotide microarray analysis. J. Biol. Chem. 278, 12029–12038 (2003)

    Article  CAS  Google Scholar 

  10. Piantadosi, C. A. et al. Heme oxygenase-1 couples activation of mitochondrial biogenesis to anti-inflammatory cytokine expression. J. Biol. Chem. 286, 16374–16385 (2011)

    Article  CAS  Google Scholar 

  11. Prochaska, H. J. & Santamaria, A. B. Direct measurement of NAD(P)H:quinone reductase from cells cultured in microtiter wells: a screening assay for anticarcinogenic enzyme inducers. Anal. Biochem. 169, 328–336 (1988)

    Article  CAS  Google Scholar 

  12. Fahey, J. W., Dinkova-Kostova, A. T., Stephenson, K. K. & Talalay, P. The “Prochaska” microtiter plate bioassay for inducers of NQO1. Methods Enzymol. 382, 243–258 (2004)

    Article  CAS  Google Scholar 

  13. Dinkova-Kostova, A. T. et al. Direct evidence that sulfhydryl groups of Keap1 are the sensors regulating induction of phase 2 enzymes that protect against carcinogens and oxidants. Proc. Natl Acad. Sci. USA 99, 11908–11913 (2002)

    Article  CAS  ADS  Google Scholar 

  14. McMahon, M., Lamont, D. J., Beattie, K. A. & Hayes, J. D. Keap1 perceives stress via three sensors for the endogenous signaling molecules nitric oxide, zinc, and alkenals. Proc. Natl Acad. Sci. USA 107, 18838–18843 (2010)

    Article  CAS  ADS  Google Scholar 

  15. Dinkova-Kostova, A. T., Kostov, R. V. & Canning, P. Keap1, the cysteine-based mammalian intracellular sensor for electrophiles and oxidants. Arch. Biochem. Biophys. 617, 84–93 (2017)

    Article  CAS  Google Scholar 

  16. Zhang, D. D. & Hannink, M. Distinct cysteine residues in Keap1 are required for Keap1-dependent ubiquitination of Nrf2 and for stabilization of Nrf2 by chemopreventive agents and oxidative stress. Mol. Cell. Biol. 23, 8137–8151 (2003)

    Article  CAS  Google Scholar 

  17. Linker, R. A. et al. Fumaric acid esters exert neuroprotective effects in neuroinflammation via activation of the Nrf2 antioxidant pathway. Brain 134, 678–692 (2011)

    Article  Google Scholar 

  18. Tallam, A. et al. Gene regulatory network inference of immunoresponsive gene 1 (IRG1) identifies interferon regulatory factor 1 (IRF1) as its transcriptional regulator in mammalian macrophages. PLoS One 11, e0149050 (2016)

    Article  Google Scholar 

  19. Naujoks, J. et al. IFNs modify the proteome of legionella-containing vacuoles and restrict infection via IRG1-derived itaconic acid. PLoS Pathog. 12, e1005408 (2016)

    Article  Google Scholar 

  20. Guarda, G. et al. Type I interferon inhibits interleukin-1 production and inflammasome activation. Immunity 34, 213–223 (2011)

    Article  CAS  Google Scholar 

  21. Thimmulappa, R. K. et al. Nrf2 is a critical regulator of the innate immune response and survival during experimental sepsis. J. Clin. Invest. 116, 984–995 (2006)

    Article  CAS  Google Scholar 

  22. Cordes, T. et al. Immunoresponsive gene 1 and itaconate inhibit succinate dehydrogenase to modulate intracellular succinate levels. J. Biol. Chem. 291, 14274–14284 (2016)

    Article  CAS  Google Scholar 

  23. Freigang, S. et al. Nrf2 is essential for cholesterol crystal-induced inflammasome activation and exacerbation of atherosclerosis. Eur. J. Immunol. 41, 2040–2051 (2011)

    Article  CAS  Google Scholar 

  24. Dinarello, C. A. Interleukin-1 in the pathogenesis and treatment of inflammatory diseases. Blood 117, 3720–3732 (2011)

    Article  CAS  Google Scholar 

  25. Shen, H. et al. The human knockout gene CLYBL connects itaconate to vitamin B12 . Cell 171, 771–782 (2017)

    Article  CAS  Google Scholar 

  26. Chappell, J. B. & Hansford, R. V. A. Subcellular Components: Preparation and Fractionation 2nd edn (Butterworth, 1972)

  27. Bridges, H. R., Mohammed, K., Harbour, M. E. & Hirst, J. Subunit NDUFV3 is present in two distinct isoforms in mammalian complex I. Biochim. Biophys. Acta 1858, 197–207 (2017)

    Article  CAS  Google Scholar 

  28. Akerboom, T. P. & Sies, H. Assay of glutathione, glutathione disulfide, and glutathione mixed disulfides in biological samples. Methods Enzymol. 77, 373–382 (1981)

    Article  CAS  Google Scholar 

  29. Booty, L. M. et al. The mitochondrial dicarboxylate and 2-oxoglutarate carriers do not transport glutathione. FEBS Lett. 589, 621–628 (2015)

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank M. McMahon and J. D. Hayes for plasmids, and Cancer Research UK (C20953/A18644) and the BBSRC (BB/L01923X/1) for financial support for ATDK. This work was supported by a Wellcome Trust Investigator award to R.C.H. (110158/Z/15/Z), a grant to M.P.M. from the Medical Research Council UK (MC_U105663142), a Wellcome Trust Investigator award to MPM (110159/Z/15/Z), and a grant to E.R.S.K. and M.S.K. from the Medical Research Council UK (MC_U105663139). B.M.K. and R.F. are supported by the Kennedy Trust Fund. We acknowledge Metabolon for their assistance with the metabolic work and analysis. The O’Neill laboratory acknowledges the following grant support: European Research Council (ECFP7-ERC-MICROINNATE), Science Foundation Ireland Investigator Award (SFI 12/IA/1531), GlaxoSmithKline Visiting Scientist Programme and The Wellcome Trust (oneill-wellcometrust-metabolic, grant number 205455). E.T.C. is supported by the Claudia Adams Barr Program.

Author information

Authors and Affiliations

Authors

Contributions

E.L.M. and D.G.R. designed and performed experiments and analysed the data. E.L.M. wrote the manuscript with assistance from all other authors. D.M., M.M.H., M.C.R. and A.F.M. performed in vitro experiments using OI. R.G.C., D.C.S., A.S.H.C. and C.F. assisted with the metabolomics analysis. Z.Z., P.G.F. and E.H. assisted with the in vivo mouse LPS trials. S.T.C. and R.C.H. were responsible for the design and synthesis of octyl esters. H.A.P., E.R.S.K., M.S.K. and L.M.B. assessed the effect of OI and itaconate on mitochondrial parameters and itaconate transport. D.D., M.H. and A.T.D.-K. performed the NQO1 assay and KEAP1 wild-type and Cys151Ser mutant experiments. J.F.M., R.F., B.M.K., E.T.C., M.P.J. and J.S. assisted with mass spectrometry experiments. L.K.M. and G.B. provided guidance and advice. E.V.K., P.J.M. and M.L.J.A. assisted with experiments in Nrf2-deficient mice. L.A.O’N. conceived ideas and oversaw the research programme. M.P.M. provided advice, reagents and oversaw a portion of the work.

Corresponding author

Correspondence to Luke A. O’Neill.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Additional information

Reviewer Information Nature thanks N. S. Chandel, R. Rossignol, S. Werner and the other anonymous reviewer(s) for their contribution to the peer review of this work.

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 Figure 1 The effect of itaconate on complex II activity.

a, Complex II and III activity in bovine heart mitochondrial membranes incubated with succinate plus malonate or itaconate (n = 3 independent experiments). b, Effect of malonate or itaconate on the oxygen consumption rate (OCR) of rat liver mitochondria in the presence of succinate (1 mM) and FCCP (1 μM; n = 3 independent experiments). c, d, Complex II and III activity in bovine heart mitochondrial membranes incubated with itaconate (IA; 1 mM unless indicated), with subsequent removal and addition of succinate (1 mM; n = 3 independent experiments) (see Methods for further details). Data are mean ± s.e.m. P values calculated using one or two-way ANOVA.

Extended Data Figure 2 DMI activates Nrf2 and limits cytokine production.

a, c, LPS (100 ng ml−1)-induced Nrf2 (a, 24 h) and HMOX1 (c, 6 h) protein expression with or without the itaconate derivative DMI. b, Nrf2-dependent mRNA expression after treatment with LPS (6 h) and DMI where indicated (n = 3). d, Reduced glutathione (GSH) and oxidized glutathione (GSSG) levels after treatment with LPS and DMI (n = 5). e, f, LPS (24 h)-induced Il1b mRNA (e), IL-1β and HIF-1α protein (f) expression in mouse macrophages with or without DMI (n = 3). Data are mean ± s.e.m. P values calculated using one-way ANOVA. Blots are representative of three independent experiments. For gel source data, see Supplementary Fig. 1.

Extended Data Figure 3 OI is the best tool to assess itaconate-dependent Nrf2 activity.

a, Reactivity of DMI, itaconate and OI with thiols. b, c, Itaconate ester reactivity with GSH and glutathione-S-transferase (GST) as detailed in the Methods (n = 3). d, Itaconate levels in mouse C2C12 cells plus itaconate esters (n = 3). MI, 4-methyl itaconate. e, i, Itaconate (e) or GSH (i) levels plus LPS (6 h) and OI as indicated (n = 5). f, NQO1 activity in mouse Hepa1c1c7 cells treated with DMI or OI (48 h) and GSH (n = 8). g, h, Metabolic intermediates in GSH synthesis (h, average of five biological replicates). i, GSH levels after treatment with LPS (6 h) and/or OI (n = 5). j, GSH/GSSG ratio after treatment with OI (2 h) and H2O2 (100 μM, 24 h; n = 3) as indicated. k, HMOX1 protein levels after treatment with OI and/or H2O2 (24 h). l, Nrf2, HMOX1 and IL-1β protein levels in BMDMs pre-treated with OI, 4-octyl 2-methylsuccinate (OMS) or octyl succinate (OS), all 125 μM for 3 h with or without LPS (6 h). m, LPS-induced Nrf2 (24 h) and HMOX1 (6 h) protein expression with or without dimethyl malonate (DMM). Data are mean ± s.e.m. P values calculated using one- or two-way ANOVA. Blots are representative of three independent experiments. For gel source data, see Supplementary Fig. 1.

Extended Data Figure 4 Itaconate is transported by the mitochondrial oxoglutarate, dicarboxylate and citrate carriers.

a, Itaconate uptake into vesicles of Lactococcus lactis membranes expressing the indicated carriers loaded with itaconate (1 mM), and transport initiated by the addition of [3H]itaconate (1 μM). b, Initial transport rates of each carrier with either canonical substrate (homo-exchange) or canonical substrate/itaconate (hetero-exchange). n = 4 independent experiments; data are mean ± s.d. P values calculated using two-tailed Student’s t-test.

Extended Data Figure 5 KEAP1 is alkylated by OI on major redox sensing cysteine residues.

a, Modification of cysteine by fumarate or itaconate. Tandem mass spectrometry spectrum of KEAP1 Cys257 (b), Cys257 (c) and Cys288 (d) peptides, indicating alkylation of these sites after OI treatment (left) but not in the corresponding carbamidomethylated (CAM) peptides (right). e, f, LDHA Cys84 alkylation after treatment with LPS (e, 24 h) or OI (f, 250 μM, 4 h) (n = 4). Detected N- and C-terminal fragment ions of both peptides are assigned in the spectrum and depicted as follows: b: N-terminal fragment ion; y: C-terminal fragment ion; asterisk: fragment ion minus NH3; 0 or asterisk: fragment ion minus H2O; and 2+: doubly charged fragment ion. Representative of one independent experiment.

Extended Data Figure 6 Identification of an itaconate-cysteine adduct.

ae, 13C6-glucose (ac) or 13C5-glutamine (d, e) labelling experiment tracking itaconate-cysteine adduct formation in BMDMs treated with LPS (n = 5; 24 h). Data in b and e are expressed as the percentage isotopologue of the total pool. Data in c and f represent changes in the total pool after LPS treatment. Data are mean ± s.e.m., for five replicates. P values calculated using two-way ANOVA.

Extended Data Figure 7 OI decreases LPS-induced cytokine production, extracellular acidification rate, ROS and nitric oxide.

a, Percentage cytotoxicity (LDH release) in BMDMs after treatment with LPS and OI as indicated (n = 3). b, LPS-induced extracellular acidification rate (ECAR) after treatment with OI and/or LPS as indicated, analysed on the Seahorse XF-24 in BMDMs (trace representative of three independent experiments). c, d, LPS-induced Il10 mRNA (c, 4 h) and protein (d, 24 h) and TNF protein (f; n = 7) after OI treatment as indicated (n = 3). e, Phosphorylated p65 (pp65) protein levels (a measure of NF-κΒ activity) after treatment with LPS and OI as indicated. h, Representative gating strategy for FACS analysis of ROS production in cells as treated in d (image representative of three independent experiments). i, LPS-induced NOS2 expression (n = 6), with or without OI treatment. j, LPS-induced TNF (n = 4) and IL-1β (n = 3) protein levels after OI treatment in PBMCs. k, Nrf2 and HMOX1 protein levels or Nrf2-dependent gene expression (n = 5) in peritoneal macrophages from mice (m) injected intraperitoneally with OI (50 mg kg−1, 6 h) or vehicle control. l, Serum IL-10 from mice injected intraperitoneally with vehicle control or OI (50 mg kg−1, 2 h) and LPS (2.5 mg kg−1, 2 h, n = 3 vehicle, OI; n = 15 LPS, OI plus LPS). Data are mean ± s.e.m. P values calculated using one-way ANOVA. Blots are representative of three independent experiments. For gel source data, see Supplementary Fig. 1.

Source data

Extended Data Figure 8 The effects of OI on cytokine production are Nrf2-dependent.

ae, Nrf2, HMOX1 and IL-1β protein levels (a, c, d) and Il1b mRNA expression (b, e) in mouse BMDMs transfected with two different Nrf2 siRNAs (50 nM) compared with a non-silencing scrambled control siRNA plus LPS (6 h; ac, e; 24 h; d) and/or OI (n = 6). NT, non-transfected. f, Il1b mRNA expression in wild-type and Nrf2-knockout BMDMs treated with LPS (24 h; WT n = 2, Nrf2 KO n = 4) and/or OI. gk, Il1b (g) and Nos2 (j) mRNA, and IL-1β (h), IL-10 (i), TNF and nitrite (k) production with or without LPS (24 h), diethyl maleate (DEM; 100 μM) or 15-deoxy-Δ12,14-prostaglandin J2 (J2; 5 μM) pre-treatment for 3 h (n = 3). Data are mean ± s.e.m. P values calculated using one-way ANOVA. Blots are representative of three independent experiments. For gel source data, see Supplementary Fig. 1.

Extended Data Figure 9 An Nrf2-dependent feedback loop exists between itaconate and IFN-β.

a, Metabolite levels after treatment with IFN-β (1,000 U ml−1; 27 h; n = 5). b, c, Isg20 and Irf5 mRNA expression in BMDMs treated with LPS (b) or poly(I:C) (c, 40 μg ml−1; 24 h) and/or OI (n = 6). d, Il10 mRNA (n = 3) and IL-10 protein (n = 5) expression after treatment with LPS for 4 h (left) or 24 h (right) and/or IFN-β treatment (1,000 U ml−1) for 3 h. e, Isg20 expression in BMDMs transfected with two different Nrf2 siRNAs (50 nM) compared with non-silencing control plus LPS (6 h) and/or OI (n = 6). f, Isg20 mRNA expression in wild-type (n = 2) and Nrf2-knockout (n = 4) BMDMs plus LPS (6 h) and/or OI. g, Isg20 mRNA expression after pre-treatment with LPS (24 h) and/or diethyl maleate (100 μM) or 15-deoxy-Δ12,14-prostaglandin J2 (5 μM) for 3 h (n = 3). Data are mean ± s.e.m. P values calculated using one-way ANOVA.

Extended Data Table 1 Mass spectrometry analysis of itaconate-induced cysteine alkylation

Supplementary information

Life Sciences Reporting Summary (PDF 132 kb)

Supplementary Information

This file contains Supplementary Figure 1, the uncropped scans with size marker indications, Supplementary Methods and Supplementary References. (PDF 1617 kb)

PowerPoint slides

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mills, E., Ryan, D., Prag, H. et al. Itaconate is an anti-inflammatory metabolite that activates Nrf2 via alkylation of KEAP1. Nature 556, 113–117 (2018). https://doi.org/10.1038/nature25986

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature25986

This article is cited by

Comments

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

Search

Quick links

Nature Briefing: 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