Skip to main content
Advertisement
  • Loading metrics

Itaconate or how I learned to stop avoiding the study of immunometabolism

Introduction to itaconate

Itaconate is a mitochondrial metabolite, produced in high amounts by macrophages and monocytes of mice and humans upon activation by several inflammatory stimuli. In 2013, the identity of the itaconate-producing enzyme was revealed as aconitate decarboxylase 1, encoded by the gene Acod1 (previously known as immune-responsive gene 1, Irg1). This clear linkage of production of a mitochondrial metabolite in response to inflammatory signalling immediately raised a flurry of interest. Indeed, itaconate has significant immunomodulatory properties and may pave the way for new immunomodulatory drugs. Presented here are several aspects of itaconate biology and a discussion of future research avenues.

How is itaconate synthesised and transported?

Upon inflammatory stimuli, including microbial components like lipopolysaccharide (LPS) and fungal cell wall sugars as well as several interferon cytokines, macrophages up-regulate Acod1, the gene encoding aconitate decarboxylase 1 (or cis-aconitate decarboxylase). This mitochondrial enzyme (Fig 1) uses cis-aconitate from the citric acid cycle to synthesise itaconate. Following synthesis in the mitochondria, humans and other mammals can detoxify this metabolite with mitochondrial enzymes that convert itaconate into acetyl-CoA and pyruvate [1], but this metabolite does accumulate in high amounts for several hours after inflammatory stimuli in vitro [2]. Current evidence suggests that a pool of itaconate is exported to the cytosol, presumably by mitochondrial 2-oxoglutarate/malate carrier protein, although transporter specificity and redundancy are yet to be determined [3]. Itaconate is likely transported from the cytosol into the phagosome: Using gene-engineered bacterial sensors, itaconate was detected in the phagosomal compartment [4], presumably to exert antimicrobial activity. Similarly, addition of exogenous itaconate reverts many of the consequences of genetic deletion of Acod1, suggesting that transport into cells is efficient. A more detailed knowledge of itaconate transport within intracellular compartments and into neighbouring cells is urgently needed, particularly when one considers that itaconate may affect neighbouring cells in tissues (discussed below).

thumbnail
Fig 1. Itaconate modifies immune responses by controlling metabolism and posttranscriptionally regulating immune cascades, which is posited to prevent excessive tissue damage.

IFNβ, interferon beta; IFNγ, interferon gamma; IL, interleukin; Acod1/Irg1: aconitate decarboxylase 1/immune-responsive gene 1; KEAP1, Kelch-like ECH-associated protein 1; LPS, lipopolysaccharide; LTx, leukotriene; MHC, major histocompatibility complex; mtROS, mitochondria-derived reactive oxygen species; NRF2, nuclear factor erythroid 2–related factor 2; OxPhos, oxidative phosphorylation; PG, prostaglandin; SDH/CII, succinate dehydrogenase and complex II of the respiratory chain; TCA cycle, tricarboxylic acid cycle.

https://doi.org/10.1371/journal.ppat.1010361.g001

Is itaconate antimicrobial?

Toxicity of itaconate to microbial pathogens is only observed at millimolar concentrations since successful pathogens have evolved to detoxify, tolerate, and even commandeer itaconate for their benefit.

The role of itaconate in microbial lifestyle and microbe’s response to stress is unknown. Production of itaconate is widespread in nature, produced by several fungi, including some Aspergillus species, and bacteria, such as Bacillus subtilis [5], giving these and other microbes ample opportunity to evolve the capacity to detoxify itaconate. An estimated 11% of all bacteria, including human pathogens such as Yersinia pestis and Pseudomonas aeruginosa, possess homologues of itaconate detoxifying genes, with conserved mechanisms producing acetyl-CoA and pyruvate [6,7]. Itaconate can be used as fuel by microbes, and some pathogens have even evolved to take advantage of itaconate production by the host. For example, P. aeruginosa isolates from cystic fibrosis patients are more likely to preferentially use itaconate as a carbon source [8].

Itaconate may synergise with other antimicrobial molecules and nutritional restrictions in the host phagosomal milieu to exert some antimicrobial effects. A very recent report showed that itaconate antimicrobial activity is potentiated by low pH, suggesting that within the phagosome, itaconate may become a significant microbicide [9]. However, the current data support that direct antimicrobial effects are not a major function of immune-derived itaconate.

Still, genetic deletion of Acod1 increases mortality of mice infected by Mycobacterium tuberculosis and increases bacterial burden in mice infected with Brucella and Salmonella [1013], likely due to a combination of direct and indirect effects in the host and in the pathogen. This model of coordinated action of itaconate with other immunoregulatory and antimicrobial factors is supported by findings that the combined actions of Acod1 and 5 other immune genes, Nos2, Cybb, Irgm1, Irgm3, and Casp4, coordinate interferon gamma (IFNγ)-induced killing of Legionella in bone marrow–derived macrophages [14]. Given that both itaconate and nitric oxide (NO) [15,16] have important immunoregulatory effects on the host cell, including at the metabolic level, it is complex to dissect the immunomodulatory versus direct antimicrobial effects and requires the capability to precisely and reliably manipulate targets on each partner of the host–pathogen interaction.

How does itaconate act as an immunomodulator?

The general framework is that itaconate acts as a negative regulator of innate immunity to limit host tissue damage, with some notable exceptions. The detailed molecular mechanisms of how itaconate exerts these effects are a very active area of research, with multiple mechanisms uncovered. Early on, a zebrafish (zebra danio) model of Salmonella infection showed the homologue of Acod1 to be induced in macrophage lineage cells, in a manner dependent on the Jak/Stat pathway and glucocorticoid pathway [17]. Expression of the zebrafish (zebra danio) homologue of Acod1 leads to mitochondrial reactive oxygen species (ROS) production in this model.

First, itaconate acts on mitochondria by competitive inhibition of succinate dehydrogenase (SDH) [18]. SDH is part of both the tricarboxylic acid cycle and the respiratory chain (complex II); thus, itaconate simultaneously interrupts the tricarboxylic acid cycle and reduces SDH-dependent oxygen consumption [16,19]. This results in succinate accumulation and a direct reduction of mitochondrial ROS (in contrast to the zebrafish (zebra danio) model). From the mitochondrial matrix, itaconate is transported into the cytosol where it can regulate metabolism at the glycolysis step, inhibiting key enzymes in this pathway to decrease glycolysis [20]. Thus, itaconate mediates several of the metabolic changes during inflammation; these direct effects are observed in most in vitro inflammatory models. Whether this occurs in vivo, and how it may be influenced by tissue microenvironment during acute inflammation and resolution stages, is still to be fully demonstrated: In a mouse model of pulmonary fibrosis, tissue-resident macrophages showed decreased oxygen consumption rate (OCR) in the absence of itaconate [21], in contrast to what is predicted by in vitro models.

Once in the cytosol (Fig 1), itaconate acts by directly modifying proteins via alkylation of cysteines [3]. In particular, alkylation of Kelch-like ECH-associated protein 1 (KEAP1) releases nuclear factor erythroid 2–related factor 2 (NRF2) [3], a potent antioxidant regulator, which then activates a second pathway to reduce cellular ROS.

Itaconate shows a lasting anti-inflammatory effect after LPS stimulation, generally decreasing levels of inflammatory cytokines interleukin (IL)-6, IL-1β, and via a negative feedback loop with interferon beta (IFNβ) [3]. Itaconate controls monocyte priming and immune paralysis after LPS treatment through an SDH-dependent mechanism in a model of human endotoxemia [22]. In other models, such as delayed inflammasome priming in murine bone marrow–derived macrophages, itaconate controls inflammasome activation via a novel posttranslational modification: itaconation [23]. Additional regulatory mechanisms of itaconate are still being uncovered. For example, itaconate controls the number of lipids bodies of macrophages after mycobacteria challenge [11]. Control of lipid bodies and lipid metabolism is likely a major regulatory pathway of itaconate [24,25], since lipid bodies are precursors for important inflammatory molecules such as prostaglandin and leukotrienes [26].

How does itaconate regulate immunity at the tissue level?

Itaconate produced by myeloid cells can directly affect neighbouring cells. In a murine model of bleomycin-induced pulmonary fibrosis, itaconate produced by alveolar macrophages improved the healing pattern of lung fibroblasts; exogenous addition of itaconate influenced fibroblast healing pattern to ameliorate lung function [21]. Itaconate may also be produced by cells other than myeloid cells. Neurons were found to produce itaconate after challenged with Zika virus in vitro [27]. Curiously, neurons are able to take up exogenous itaconate, and itaconate application improved neurological function upon reperfusion injury [28]. Thus, itaconate can control tissue function by acting on nonmyeloid cells.

In mouse models of mycobacterial infections, deletion of Acod1 leads to differential recruitment of immune cells in the lung, such as an increase in neutrophils [10] and in lymphocytes [11]. However, it is still unclear which immune signal is influencing this differential recruitment and whether itaconate may influence the development of adaptive immunity. Thus far, a single study observed that itaconate did not affect protective immunity by the Bacillus Calmette–Guérin (BCG) vaccine strain [11]. This observation fits current models at a conceptual level, as an attenuated strain BCG vaccine strain will not trigger excessive inflammation, but it is very intriguing at a mechanistic level, i.e., at which point is there enough inflammation to activate itaconate and through which cellular mechanisms? Is it possible that in vivo there is a need for tissue damage, and danger associated molecular patterns, to trigger production of itaconate? Thus, further studies understanding impact of itaconate in tissue homeostasis and influence on adaptive immunity are urgently required.

Are analogues of itaconate potential immunomodulatory therapies?

In human studies, low levels of itaconate in plasma are associated with excessive inflammation in rheumatoid arthritis [28] and during Coronavirus Disease 2019 (COVID-19) [29], which strongly supports that supplement of itaconate and its analogues would be useful in clinic.

Currently, several analogues of itaconate show potent immunomodulatory properties. However, while these mimic several of itaconate functions, they show important differences from endogenously produced itaconate [25,29,30]. For example, analogues of itaconate prevents cycloxigenase-2 expression and production of several prostaglandin species in response to a Toll-like receptor (TLR)1/2 agonist Pam3CSK4, a possibly clinically useful anti-inflammatory effect, an effect not replicated by endogenous itaconate [25]. This is attributed to different reactivity of these analogues to act as cysteine modifiers and electrophilic molecules. Itaconate analogues showed beneficial effects in mouse models of psoriasis [31] and others autoimmune diseases. Interestingly, both dimethyl fumarate, a compound in clinical use to control psoriasis, and itaconate analogues were able to decrease prostaglandin expression in macrophages [25], suggesting some mechanistic and clinical overlap of itaconate analogues to existing therapies.

Overall, itaconate, and the immunobiology under its control, is revealing a wealth of knowledge on delicate equilibrium required of a successful immune response. While considerable work is still needed to fully understand these cascades, this knowledge holds great potential to improve our management of infectious and immune diseases.

References

  1. 1. Shen H, Campanello GC, Flicker D, Grabarek Z, Hu J, Luo C, et al. The Human Knockout Gene CLYBL Connects Itaconate to Vitamin B12. Cell. 2017;171:771–82.e11. pmid:29056341
  2. 2. Michelucci A, Cordes T, Ghelfi J, Pailot A, Reiling N, Goldmann O, et al. Immune-responsive gene 1 protein links metabolism to immunity by catalyzing itaconic acid production. Proc Natl Acad Sci U S A. 2013;110:7820–5. pmid:23610393
  3. 3. Mills EL, Ryan DG, Prag HA, Dikovskaya D, Menon D, Zaslona Z, et al. Itaconate is an anti-inflammatory metabolite that activates Nrf2 via alkylation of KEAP1. Nature. 2018;556:113–7. pmid:29590092
  4. 4. Naujoks J, Tabeling C, Dill BD, Hoffmann C, Brown AS, Kunze M, et al. IFNs Modify the Proteome of Legionella-Containing Vacuoles and Restrict Infection Via IRG1-Derived Itaconic Acid. PLoS Pathog. 2016;12:e1005408. pmid:26829557
  5. 5. Chun HL, Lee SY, Lee SH, Lee CS, Park HH. Enzymatic reaction mechanism of cis-aconitate decarboxylase based on the crystal structure of IRG1 from Bacillus subtilis. Sci Rep. 2020;10:11305. pmid:32647315
  6. 6. Sasikaran J, Ziemski M, Zadora PK, Fleig A, Berg IA. Bacterial itaconate degradation promotes pathogenicity. Nat Chem Biol. 2014;10:371–7. pmid:24657929
  7. 7. Wang H, Fedorov AA, Fedorov EV, Hunt DM, Rodgers A, Douglas HL, et al. An essential bifunctional enzyme in Mycobacterium tuberculosis for itaconate dissimilation and leucine catabolism. Proc Natl Acad Sci U S A. 2019;116:15907–13. pmid:31320588
  8. 8. Riquelme SA, Liimatta K, Wong Fok Lung T, Fields B, Ahn D, Chen D, et al. Pseudomonas aeruginosa Utilizes Host-Derived Itaconate to Redirect Its Metabolism to Promote Biofilm Formation. Cell Metab. 2020;31:1091–106.e6. pmid:32428444
  9. 9. Duncan D, Lupien A, Behr MA, Auclair K. Effect of pH on the antimicrobial activity of the macrophage metabolite itaconate. Microbiology. 2021;167. pmid:34020726
  10. 10. Nair S, Huynh JP, Lampropoulou V, Loginicheva E, Esaulova E, Gounder AP, et al. Irg1 expression in myeloid cells prevents immunopathology during M. tuberculosis infection. J Exp Med. 2018;215:1035–45. pmid:29511063
  11. 11. Hoffmann E, Machelart A, Belhaouane I, Deboosere N, Pauwels A-M, Saint-André J-P, et al. IRG1 controls immunometabolic host response and restricts intracellular Mycobacterium tuberculosis infection. bioRxiv. 2019 Sep.
  12. 12. Meixin C, Hui S, Maikel B, Lin S, Shu-Jung C, Weiwei W, et al. Itaconate is an effector of a Rab GTPase cell-autonomous host defense pathway against Salmonella. Science. 2020;369:450–5. pmid:32703879
  13. 13. Demars A, Vitali A, Comein A, Carlier E, Azouz A, Goriely S, et al. Aconitate decarboxylase 1 participates in the control of pulmonary Brucella infection in mice. Tsolis RM, editor. PLoS Pathog. 2021;17:e1009887. pmid:34525130
  14. 14. Price JV, Russo D, Ji DX, Chavez RA, DiPeso L, Lee AY-F, et al. IRG1 and Inducible Nitric Oxide Synthase Act Redundantly with Other Interferon-Gamma-Induced Factors To Restrict Intracellular Replication of Legionella pneumophila. mBio. 2019;10:e02629–19. pmid:31719183
  15. 15. Brown GC, Foxwell N, Moncada S. Transcellular regulation of cell respiration by nitric oxide generated by activated macrophages. FEBS Lett. 1998;439:321–4. pmid:9845346
  16. 16. Lampropoulou V, Sergushichev A, Bambouskova M, Nair S, Vincent EE, Loginicheva E, et al. Itaconate Links Inhibition of Succinate Dehydrogenase with Macrophage Metabolic Remodeling and Regulation of Inflammation. Cell Metab. 2016;24:158–66. pmid:27374498
  17. 17. Hall CJ, Boyle RH, Astin JW, Flores MV, Oehlers SH, Sanderson LE, et al. Immunoresponsive Gene 1 Augments Bactericidal Activity of Macrophage-Lineage Cells by Regulating β-Oxidation-Dependent Mitochondrial ROS Production. Cell Metab. 2013;18:265–78. pmid:23931757
  18. 18. Cordes T, Metallo CM. Itaconate Alters Succinate and Coenzyme A Metabolism via Inhibition of Mitochondrial Complex II and Methylmalonyl-CoA Mutase. Metabolites. 2021;11:117. pmid:33670656
  19. 19. Németh B, Doczi J, Csete D, Kacso G, Ravasz D, Adams D, et al. Abolition of mitochondrial substrate-level phosphorylation by itaconic acid produced by LPS-induced Irg1expression in cells of murine macrophage lineage. FASEB J. 2016;30:286–300. pmid:26358042
  20. 20. Sakai A, Kusumoto A, Kiso Y, Furuya E. Itaconate reduces visceral fat by inhibiting fructose 2,6-bisphosphate synthesis in rat liver. Nutrition. 2004;20:997–1002. pmid:15561490
  21. 21. Ogger PP, Albers GJ, Hewitt RJ, O’Sullivan BJ, Powell JE, Calamita E, et al. Itaconate controls the severity of pulmonary fibrosis. Sci Immunol. 2020;5:eabc1884. pmid:33097591
  22. 22. Domínguez-Andrés J, Novakovic B, Li Y, Scicluna BP, Gresnigt MS, Arts RJW, et al. The Itaconate Pathway Is a Central Regulatory Node Linking Innate Immune Tolerance and Trained Immunity. Cell Metab. 2019;29:211–20.e5. pmid:30293776
  23. 23. Bambouskova M, Potuckova L, Paulenda T, Kerndl M, Mogilenko DA, Lizotte K, et al. Itaconate confers tolerance to late NLRP3 inflammasome activation. Cell Rep. 2021;34:108756. pmid:33691097
  24. 24. Verberk SGS, Kuiper KL, Lauterbach MA, Latz E, Van den Bossche J. The multifaceted therapeutic value of targeting ATP-citrate lyase in atherosclerosis. Trends Mol Med. 2021;27:1095–105. pmid:34635427
  25. 25. Diskin C, Zotta A, Corcoran SE, Tyrrell VJ, Zaslona Z, O’Donnell VB, et al. 4-Octyl-Itaconate and Dimethyl Fumarate Inhibit COX2 Expression and Prostaglandin Production in Macrophages. J Immunol. 2021;207:2561–9. pmid:34635585
  26. 26. Melo RCN, D’Avila H, Wan H-C, Bozza PT, Dvorak AM, Weller PF. Lipid Bodies in Inflammatory Cells: Structure, Function, and Current Imaging Techniques. J Histochem Cytochem. 2011;59:540–56. pmid:21430261
  27. 27. Daniels BP, Kofman SB, Smith JR, Norris GT, Snyder AG, Kolb JP, et al. The Nucleotide Sensor ZBP1 and Kinase RIPK3 Induce the Enzyme IRG1 to Promote an Antiviral Metabolic State in Neurons. Immunity. 2019;50:64–76.e4. pmid:30635240
  28. 28. Cordes T. Itaconate modulates tricarboxylic acid and redox metabolism to mitigate reperfusion injury. Mol Metab. 2020;32:122–35. pmid:32029222
  29. 29. ElAzzouny M, Tom CTMB, Evans CR, Olson LL, Tanga MJ, Gallagher KA, et al. Dimethyl Itaconate Is Not Metabolized into Itaconate Intracellularly. J Biol Chem. 2017;292:4766–9. pmid:28188288
  30. 30. Swain A, Bambouskova M, Kim H, Andhey PS, Duncan D, Auclair K, et al. Comparative evaluation of itaconate and its derivatives reveals divergent inflammasome and type I interferon regulation in macrophages. Nat Metab. 2020;2:594–602. pmid:32694786
  31. 31. Bambouskova M, Gorvel L, Lampropoulou V, Sergushichev A, Loginicheva E, Johnson K, et al. Electrophilic properties of itaconate and derivatives regulate the IκBζ–ATF3 inflammatory axis. Nature. 2018;556:501–4. pmid:29670287