Skip to main content

The role and therapeutic potential of itaconate in lung disease

Abstract

Lung diseases triggered by endogenous or exogenous factors have become a major concern, with high morbidity and mortality rates, especially after the coronavirus disease 2019 (COVID-19) pandemic. Inflammation and an over-activated immune system can lead to a cytokine cascade, resulting in lung dysfunction and injury. Itaconate, a metabolite produced by macrophages, has been reported as an effective anti-inflammatory and anti-oxidative stress agent with significant potential in regulating immunometabolism. As a naturally occurring metabolite in immune cells, itaconate has been identified as a potential therapeutic target in lung diseases through its role in regulating inflammation and immunometabolism. This review focuses on the origin, regulation, and function of itaconate in lung diseases, and briefly discusses its therapeutic potential.

Introduction

Lung diseases remain prevalent clinical conditions with high morbidity and mortality rates. However, targeted drugs are limited to a few diseases, and for most lung conditions, medical management primarily aims at symptom relief and control of disease progression [1,2,3]. The development of many lung diseases can be attributed to tissue inflammation and oxidative stress injury, whether due to infections or malignancies [2, 4]. An impaired balance between inflammation and anti-inflammation, along with oxidative stress, has been identified as a hallmark in most pulmonary diseases, including lung cancer, acute and chronic lung injury induced by infection or physical/chemical exposure, and pulmonary fibrosis. Despite the varying pathologies, the immune system plays a significant role in these lung diseases. The immune response in lung disease is characterized by altered metabolism in different cellular components. For example, glycolysis in typical M1 macrophages has been demonstrated to be an important initiator and promoter of inflammation.

Given the recognition of the immune system as a key regulator in lung disease, immune metabolism reprogramming has emerged as a potential therapeutic target. The altered metabolism in immune cells, as a reaction and adaptation to both external and internal stimuli, characterizes their functional phenotype. For example, the Warburg effect, which involves increased glucose consumption for the production of biomacromolecule precursors and nicotinamide adenine dinucleotide phosphate (NADPH), occurs during acute stimulation, ensuring cellular survival and proliferation under external stress exposure. In contrast, fatty acid oxidation and the classical tricarboxylic acid (TCA) cycle maintain metabolism in quiescent cells. Interestingly, both metabolites and enzymes involved in these processes can regulate immunometabolism and immune function in return, such as nitric oxide (NO), arginine, and glutamine, revealing a complex interaction. Consequently, efforts have been made to identify the therapeutic potential of immunometabolism regulation in lung disease. For instance, the protective effect of dexamethasone in asthma is associated with downregulation of pyruvate dehydrogenase kinase, glutaminase, and fatty acid synthase [5]. Potential immunometabolic targets in some lung diseases have been previously reported, including the NOD-like receptor protein 3 (NLRP3) inflammasome, the cGAS-STING axis, and beta-hydroxybutyrate [6]. As a metabolite from TCA cycle, itaconate has emerged as a novel and potential metabolic target.

With the onset of the coronavirus disease 2019 (COVID-19) pandemic, lung diseases such as acute lung injury and pulmonary fibrosis have garnered increased attention. Lung diseases are among the leading causes of death globally and significantly impact the quality of life. Itaconate, a macrophage metabolite that has emerged in recent years, has been shown to possess anti-inflammatory and anti-oxidative stress effects, as well as effective immunometabolism regulating functions. This review focuses on the precise anti-inflammatory and anti-oxidative stress functions of itaconate and its exogenous derivatives in various lung diseases, along with their potential mechanisms. By summarizing recent studies primarily on pulmonary inflammatory diseases, it is demonstrated that both itaconate and its exogenous derivatives have anti-inflammatory and antioxidant effects, offering significant pharmaceutical potential.

The origin and exploration of itaconate

Itaconate was first synthesized by Jean Louis Lassaigne in 1836 during his study of the thermal decomposition of citric acid [7]. Initially, it was used for industrial polymer synthesis until its role in the TCA cycle was investigated in the twentieth century [8]. While itaconate is a potential intermediate in central metabolism, it cannot support respiration in the same way as succinate or malate [8]. In 2004, Sakai and colleagues reported that itaconate induced suppressed glycolysis levels in liver cells, supporting its function as a metabolism regulator [9]. The presence of itaconate in mammalian cells was first observed in 2011 [10], and metabolite profiling revealed increased levels of itaconate in macrophages during lipopolysaccharide (LPS) stimulation [11]. In 2013, Michelucci and colleagues demonstrated that itaconate is endogenously produced in macrophages by the mammalian cis-aconitate decarboxylase (CAD), also known as immune response gene 1 protein (IRG1), which encodes aconitate decarboxylase 1 (ACOD1) [12]. Previous studies of IRG1 had identified it as one of the most highly upregulated genes under pro-inflammatory conditions, highlighting its crucial role in the immune response [13, 14]. Thus, IRG1 links cellular metabolism with immune defense by catalyzing the production of itaconate [12]. Increasing evidence from subsequent studies suggests that itaconate has important anti-inflammatory and antioxidant effects in mammals [15, 16]. Many efforts have been made to explore the mechanisms of itaconate function, and the main targets of itaconate that have been reported are summarized in Table 1. Itaconate has been reported as a succinate dehydrogenase (SDH) inhibitor [17, 18] and can inhibit mitochondrial substrate-level phosphorylation [19]. Further studies have shown that itaconate can regulate nuclear factor erythroid 2-related factor 2 (Nrf2) signaling, the IκBζ-ATF3 axis, and NLRP3 inflammasome activation [20,21,22]. Fructose-bisphosphate aldolase A, Janus kinase 1 (JAK1), TET DNA dioxygenases, and transcription factor EB (TFEB) have also been identified as itaconate targets [23,24,25,26].

Table 1 Currently reported targets and effects of itaconate

The regulation of itaconate and ACOD1

Itaconate, derived from citrate, is primarily produced by macrophages and myeloid cells under M1 polarization conditions [27]. Immune organs such as lymph nodes and spleen have the highest concentrations of itaconate [28]. The synthesis of this immune metabolite is induced by the enzyme CAD, encoded by ACOD1/IRG1, which catalyzes the decarboxylation of cis-aconitate to itaconate. This enzyme has low basal expression under normal conditions, but its expression increases upon exposure to environmental contaminants or infections [12, 29,30,31]. The anaplerotic TCA cycle in active macrophages produces high levels of itaconate [32]. As the only enzyme capable of catalyzing itaconate synthesis, IRG1’s expression is tightly regulated, though its regulatory mechanisms are still not fully understood. Interferon regulatory factor 1 is the most important transcriptional regulator of IRG1 [33]. TFEB, a lysosomal biogenesis factor activated by bacterial stimuli, drives IRG1 expression and itaconate synthesis in macrophages [34]. Itaconate in macrophages can induce TFEB nuclear translocation and activation by inducing TFEB alkylation [26]. The stimulator of interferon response cGAMP interactor 1 is also essential for LPS-induced IRG1 expression [35]. Additionally, NO and isocitrate dehydrogenase activity regulate the TCA cycle and itaconate production, thereby modulating respiratory function. A deficiency in NO in macrophages leads to increased itaconate accumulation and elevated interleukin-1β (IL-1β) level [36].

The expression of IRG1 can be regulated by several signaling pathways. Activated nuclear factor kappa-B (NF-κB) signaling can induce IRG1 expression, leading to the accumulation of itaconate in macrophages [37]. Increased IRG1 expression was also observed in A20-deficient macrophages [38]. Since the ubiquitin enzyme A20 can inhibit NF-κB, IRG1 might be upregulated by A20 through its inhibition of NF-κB signaling. The IRF9–IRG1 pathway was also previously reported, showing that decreased IRF9 levels, due to miR93, lead to the inhibition of IRG1 [39]. Other novel signaling pathways, such as the NOTCH4/GATA4/IRG1 axis, the SR-A1/STAT3/IRG1 axis, and TFEB/IRG1, have also been demonstrated as potential regulators of itaconate [34, 40, 41]. Additionally, specific knockout of peroxisome proliferator-activated receptor gamma (PPARγ) in macrophages leads to increased IRG1 expression, suggesting a potential regulatory effect of PPARγ on itaconate production [42]. IFN-I decreases itaconate production through IRG1 inhibition induced by IL-10 [43].

An interesting fact is that, although IRG1 can be upregulated by activated NF-κB signaling, the consequent accumulation of itaconate promotes Nrf2/HO-1 signaling and inhibits signal transducer and activator of transcription 3 expression, thereby suppressing NF-κB signaling [44, 45]. Additionally, reactive oxygen species (ROS) production in a murine colitis model is decreased by 4-octyl itaconate (4-OI), which inhibits the activation of MAPK/NF-κB signaling [46]. The regulation of itaconate and IRG1/ACOD1 is presented in Fig. 1.

Fig. 1
figure 1

The regulation of itaconate and IRG1. A typical upregulation of IRG1 occurs in macrophages after the LPS stimuli. The cis-aconitate are transferred and readjusted from TCA-cycle, leading to the accumulation of itaconate. The IRG1 expression can be regulated by many pathways, such as Norch/GATA4, NF-κB, STING, or phospho-STAT3; meanwhile, itaconate production can be suppressed by NO and citraconate. Pointed arrows and blunt head arrows indicate promotion and inhibition, respectively

The function of itaconate

Itaconate regulates immunometabolism and inflammation

IRG1 encodes a mitochondrial metabolic enzyme that catalyzes itaconate production. Under pathological conditions, such as LPS stimulation, the upregulation of the IRG1 enzyme stimulates the conversion of cis-aconitate to itaconate in the TCA cycle [26]. Previous studies have highlighted that itaconate functions as an important anti-inflammatory metabolite, acting as an SDH inhibitor, Nrf2 regulator, and electrophilic stress regulator [32]. Itaconate is partially upregulated by interferon-β (IFN-β) and promotes an anti-inflammatory response through Nrf2 activation and NLRP3 inhibition [16, 47]. Unmodified natural itaconate inhibits inflammasome activation and increases LPS-induced production of IFN-β in macrophages [48]. The anti-inflammatory effect of itaconate is related to SDH inhibition and changes in mitochondrial respiration in active inflammatory macrophages. The endogenous accumulation of itaconate in macrophages leads to the accumulation of succinate, a metabolite associated with the metabolic reprogramming of immune cells [17, 18]. Itaconate is metabolized into itaconyl-CoA, which modulates methylmalonyl-CoA-dependent branched-chain amino acid metabolism, altering the balance of CoAs and fatty acid diversity [49]. Additionally, itaconate stabilizes the expression of carnitine palmitoyltransferase 1A (CPT1A) by interfering with CPT1A ubiquitination, thereby enhancing lipid clearance and modulating lipid metabolism [50]. These results demonstrate that itaconate is a key regulator in TCA cycle remodeling.

4-Octyl itaconate (4-OI), a cell-permeable derivative of itaconate, blocks the activation of the NLRP3 inflammasome and IL-1β release as well as inhibits STING expression and PI3K/Akt signaling [22, 51, 52]. Additionally, itaconate derivatives can directly modify and inhibit JAK1 [25]. Another membrane-permeable itaconate derivative, dimethyl itaconate (DI), can suppress the secretion of cytokines IL-1β and C–C motif chemokine ligand 2 in epithelial cells and reduce macrophage recruitment. This leads to alleviation in ulcerative colitis in a mouse model and promotes T cell differentiation [53]. Exposure to LPS treatment upregulates IRG1 expression in macrophages, resulting in increased itaconate levels. This increase in itaconate levels may cause mitochondrial substrate-level phosphorylation loss due to potential CoA trapping [19]. The fluctuations in IRG1 and itaconate levels in macrophages might reduce the survival of infectious microbes or impair macrophage function under hypoxic stress [19]. The absence of IRG1 expression in myeloid cells contributes to the activation of NF-κB signaling, leading to excessive neutrophil recruitment and increased mortality in mice during infection. This excessive inflammation can be mitigated by exogenous itaconate supplementation [54]. Itaconate and DI can interfere with IκBζ signaling in inflammatory responses by increasing ATF3 expression [21]. Additionally, itaconate can target specific pathogens to inhibit bacterial growth [53]. By disrupting the Fe–S cluster structure within aconitase, itaconate inhibits its activity and regulates iron metabolism in macrophages [55]. Itaconate directly binds to teneleventranslocation-2, inhibiting its activity, which results in decreased 5hmC levels and protects mice from LPS-induced injury [24]. Increased itaconate levels have also been reported to inhibit TET DNA dioxygenases, thereby suppressing inflammation and reducing immune cell infiltration into tumors [37]. Structurally similar to α-ketoglutaric acid (α-KG), a dicarboxylic acid with 4- or 5-carboxylate groups, itaconate can act as an α-KG antagonist. This similarity allows ITA to selectively inhibit TET enzymes and suppress inflammatory responses. To further explore downstream pathways, Chen et al. treated TET2/TET2-mutant BMDM cells with or without 4-OI under LPS stimuli. Gene expression profiling revealed that 607 genes were downregulated in TET2-mutant cells, with these genes being enriched in those involved in innate immunity and inflammatory response. Additionally, the results suggested that IκBζ, a transcriptional regulator of selective NF-κB target genes, is encoded by Nfkbiz [24]. Despite its significant anti-inflammatory effects, high doses of itaconate can induce apoptosis and IL-1β release [56].

Itaconate also inhibits virus replication by suppressing SDH activity, which is associated with changes in cellular metabolism [57]. Moreover, itaconate can inhibit fumarate hydratase activity, leading to the accumulation of fumarate and succinate, which are crucial for intracellular energy metabolism [58]. In LPS-activated RAW264.7 cells, itaconate reduces the levels of glycolytic intermediates that are elevated in metabolically reprogrammed macrophages [59]. Itaconate inhibits fructose-bisphosphate aldolase A and glyceraldehyde 3-phosphate dehydrogenase (GAPDH), impairing glycolysis and consequently attenuating the inflammatory response in macrophages [23, 60].

Itaconate regulates oxidative stress

Itaconate and its analogs have been shown to influence the oxidation–reduction process enhance the expression of antioxidant genes, indicating its potential role in regulating oxidative stress [61, 62]. Itaconate has been reported to modulate mitochondrial ROS production in macrophages [63], and supplementation with the itaconate analog 4-OI can restore mitochondrial redox balance [64]. Upon exposure to endotoxin, IRG1 expression is markedly increased, leading to elevated A20 expression through ROS production in macrophages, which enhances endotoxin tolerance [65]. ROS production by IRG1 is regulated by the pentose phosphate pathway [66]. Itaconate can activate Nrf2 signaling [67, 68]. Itaconate alleviates hepatic ischemia–reperfusion injury by activating the Nrf2-antioxidant pathway, thereby protecting hepatocytes from oxidative stress damage [69]. Additionally, itaconate can regulate the Nrf2-mediated antioxidant response by suppressing fumarate hydratase activity [58]. In the cerebral ischemia–reperfusion model, itaconate induces a cellular antioxidant response and modulates brain redox metabolism, which protects mitochondrial function, improves brain function, and reduces mortality following reperfusion injury [70].

Itaconate regulates cell death

IRG1 is known to be associated with tumor cell proliferation, migration, and invasion, and is highly expressed in tumor cells [45]. Itaconate also increases the expression of hypoxia-inducible factor 1 targets, such as Hk2 and Vegfa [71]. In sepsis-induced mouse models, itaconate elevates Nrf2 levels and protects THP-1 cells from ferroptosis [72], while preserving mitochondrial function and inhibiting the cGAS-STING-IRF3 pathway [73]. Additionally, itaconate modulates LPS-induced pyroptosis by preventing caspase-1 activation and promoting NLRP3 inflammasome tolerance, thus mitigating tissue damage and preventing pyroptotic cell death [74]. Conversely, 4-OI has been reported to activate ferritinophagy and induce ferroptosis in retinoblastoma cells [75]. Itaconate thus appears to suppress cell death in normal cells while inducing ferroptosis in specific tumor cells through different mechanisms, highlighting its complex role in the cell cycle. Figure 2 summarized the known mechanisms by which itaconate regulates immune function, oxidative stress and cell death.

Fig. 2
figure 2

How itaconate regulates immunometabolism, inflammatory response, oxidative stress, and cell death. As a typical immune metabolite that accumulated in LPS-activated macrophages, itaconate is a crucial anti-inflammatory metabolite that acts via different manners. Nrf2-dependent anti-inflammatory function is one main target of itaconate. Itaconate can also inhibit NLRP3 activation, SDH activity, and cGAS-STING-IRF3 pathway. Itaconate interferes with CPT1A ubiquitination to stabilize CPT1A expression. Through this complex regulatory network, itaconate exquisitely models the metabolic alteration and inflammatory response to limit inflammation. Pointed arrows and blunt head arrows indicate promotion and inhibition, respectively

The role of itaconate in lung disease

Acute lung injury

Acute lung injury (ALI) is a critical clinical condition characterized by its rapid onset and high mortality rate, making it one of the leading causes of death in critically ill patients. ALI is characterized by impaired vascular endothelial and alveolar epithelial functions, leading to inflammatory infiltration, pulmonary edema, and arterial hypoxemia [76]. The progression of ALI is primarily driven by inflammation and oxidative stress. Activated platelets contribute to the pathogenesis of ALI by promoting innate immune responses, such as neutrophil recruitment and the production and secretion of proteases and toxic mediators in the lungs [77]. Concurrently, various inflammatory stimuli trigger pulmonary endothelial cells, alveolar cells, airway epithelial cells, and alveolar macrophages to produce ROS and reactive nitrogen species, which further exacerbate lung damage [78]. Currently, managing inflammation is the main therapeutic approach for ALI. Itaconate, with its anti-inflammatory and antioxidant properties, may offer potential medicinal value for treating this condition.

Itaconate exerts anti-inflammatory effects through SDH inhibition, Nrf2 activation, and reduction of pro-inflammatory cytokine production [79]. In an LPS-induced ALI model, the administration of 4-OI significantly reduces the production of pro-inflammatory mediators and ROS [72]. The protective effects of 4-OI have also been observed in two additional models: the LPS-induced murine acute kidney injury model and the microglia inflammation model [80, 81]. Similarly, in a methicillin-resistant Staphylococcus aureus-induced ALI model, 4-OI administration improved survival, mitigated pathological damage, and inhibited neutrophil infiltration [82]. Furthermore, 4-OI reduces LPS-induced NLRP3 inflammasome activation and interleukin-1β (IL-1β) secretion [83]. This effect is attributed to the alkylation-mediated dissociation of NLRP3 from NEK7 [22]. Additionally, itaconate reconstitution restores NLRP3 inflammasome tolerance in IRG1-deficient macrophages [74]. DI has been shown to inhibit LPS-induced microglia activation via the Nrf2/HO-1 signaling pathway [84]. Collectively, these studies suggest that itaconate and its analogs offer significant protective effects and have considerable clinical potential in managing inflammatory responses.

In addition to inflammatory infiltration, ALI causes redox imbalance, oxidative damage, and DNA damage, which collectively lead to lung cell death and dysfunction. Oxidative and antioxidant pathways are regulated by various external stimuli. 4-OI promotes the nuclear transport of Nrf2, a crucial endogenous antioxidant transcription factor, thereby enhancing the antioxidant response downstream of the Nrf2 signaling pathway [44]. In the context of oxidative stress in vivo, lipids are also subject to peroxidation. As previously reported, 4-OI alleviates ALI by inhibiting glutathione peroxidase 4-dependent lipid peroxidation, which is achieved through increased Nrf2 accumulation and activation [72]. Similarly, when the Nrf2 signaling pathway is blocked, the protective effects of 4-OI are significantly diminished in chondrocytes and mouse arthritis models. The overaccumulation of neutrophil extracellular traps (NETs)—which form specialized bactericidal structures following necrosis or apoptosis—can cause tissue damage. 4-OI inhibits NET formation in mouse neutrophils via the Nrf2/HO-1 pathway, thereby contributing to a favorable long-term prognosis [85]. Additionally, itaconate enhances the resistance of pulmonary microvascular endothelial cells to inflammation and mitochondrial oxidative stress by activating the Nrf2 pathway [86]. Itaconate also promotes the nuclear translocation of TFEB to regulate autophagic flux, thus reducing apoptosis [87].

The results of these studies suggest that itaconate has the potential to reduce inflammation and restore redox balance during tissue injury. Specifically, itaconate may exert its therapeutic effects through multiple pathways. For instance, reduced ROS production may influence ROS-mediated PI3K/Akt/NF-κB signaling, potentially contributing to the alleviation of LPS-induced lung inflammation by itaconate. Additionally, itaconate is known to inhibit TET enzymes, which may also partially explain its benefits in managing inflammation. Another challenge in using itaconate as a therapeutic agent for lung injury is coordinating its effects with the response to infecting pathogens. Since pathogens may also respond to changes in environmental itaconate levels, further studies are needed to determine whether pathogens adapt to exogenous itaconate supplementation and how this adaptation might affect disease progression. The antibacterial, anti-inflammatory, and antioxidant properties of itaconate result from the interaction and coordination of multiple mechanisms, reflecting its complexity and versatility as a regulator of immune metabolism rather than a mere metabolite.

Respiratory infection

As the organ responsible for gas exchange with the external environment, the lungs are naturally exposed to a wide range of potential pathogens, including bacteria, viruses, and fungi. Itaconate, as a potential immunomodulator, has been explored for its therapeutic potential in various pathogen infections. Gu et al. reported that dimethyl itaconate (DI) protects against fungal keratitis by activating the Nrf2/HO-1 signaling pathway and inhibiting the growth of Aspergillus fumigatus [88]. However, there is limited evidence that itaconate protects hosts from fungal infections. Given that fungi are a significant source of industrial itaconate [89], understanding itaconate metabolism in fungal infections is essential. In influenza A virus infection, itaconate and its analogs reduce inflammation and inhibit viral replication [90,91,92]. Notably, itaconate inhibits virus replication not through Nrf2-independent mechanisms but by directly interfering with the nuclear export of viral ribonucleoprotein complexes [91, 92]. Another area of interest is the interaction between bacteria and itaconate, as itaconate can directly impact bacterial cells. itaconate enhances NADPH oxidase activity, leading to increased ROS production and inhibition of bacterial growth [93]. The antimicrobial capacity of itaconate is also associated with its disruption of central carbon metabolism [94, 95]. Itaconate can covalently modify the active-site cysteine of isocitrate lyase, thereby inhibiting the enzyme and suppressing bacterial growth [96, 97]. However, bacteria can adapt to changes in environmental itaconate levels in several ways. For example, Staphylococcus aureus experiences ammonium starvation under itaconate stress [98], and the LysR-type transcriptional regulator RipR protects Gram-negative bacteria from itaconate challenges [99].

The novel coronavirus epidemic has resulted in over 50 million deaths and countless infections worldwide. As a disease that has garnered significant attention in recent years, numerous studies have explored the relationship between itaconate and novel coronaviruses. Severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) is a positive-stranded RNA beta-coronavirus that facilitates receptor recognition and membrane fusion through the spike glycoprotein (S protein) on the surface of the virion. This interaction with the host receptor, angiotensin-converting enzyme 2 (ACE2), mediates viral entry and can lead to severe respiratory syndrome [100]. SARS-CoV-2 drives a cytokine storm and causes immunosuppression, resulting in a hyperinflammatory state that contributes to multi-organ failure, particularly respiratory failure due to acute respiratory distress syndrome (ARDS) [101, 102]. Given that cytokine storms and oxidative stress are major contributors to ARDS during respiratory viral infections, anti-inflammatory and antioxidant therapies have emerged as effective therapeutic strategies.

Itaconate is a potential agent in the treatment of COVID-19 due to its antioxidant and immune metabolism-regulating effects. Metabolic remodeling plays a crucial role in viral infections, as host metabolism is integral to both the host immune response and viral propagation [103]. Itaconate regulates metabolic remodeling by modulating electron transport chain flux and lipid metabolism, which in turn affects macrophage activation and inflammatory responses [18, 103]. Analysis of a publicly available transcriptomic dataset from SARS-CoV-2 patient lung biopsies revealed that the Nrf2 antioxidant gene expression pathway is inhibited. David et al. found that in vitro administration of itaconate can induce cellular antiviral programs via a type I interferon-independent pathway [104]. Additionally, Nrf2 significantly inhibits SARS-CoV-2 replication by activating downstream HMOX1 to produce biliverdin, which scavenges ROS [104]. However, SARS-CoV-2 interacts with the catalytic domain of the NAD-dependent deacetylase sirtuin 1 (SIRT1) via the nonstructural viral protein NSP14, thereby inhibiting the activation of the Nrf2/HMOX1 pathway [105]. Similarly, decreased levels of Nrf2 have been observed in pediatric patients infected with SARS-CoV-2, resulting in reduced total antioxidant status and increased oxidative stress, leading to tissue damage.

Physiologically, the presence of ROS leads to the dissociation of the Kelch-like ECH-associated protein 1 (Keap1)-Nrf2 complex, allowing Nrf2 to migrate to the nucleus. This migration triggers antioxidant responses, enhancing protection against inflammation. During viral infections, such as SARS-CoV-2, inflammatory processes and oxidative stress in epithelial and endothelial cells activate the transcription factor Nrf2, which similarly protects cells from oxidative stress and inflammation. Concurrently, SARS-CoV-2 infection is associated with alterations in lipid metabolism. The Spike protein on the surface of SARS-CoV-2 impairs lipid metabolism and autophagy pathways in host cells by upregulating Nrf2, leading to increased siderosis and heightened susceptibility to lipid toxicity [106]. This spike-induced impairment in lipid metabolism can be inhibited by Nrf2 inhibitors [106]. However, Nrf2 deficiency can upregulate angiotensin-converting enzyme 2 (ACE2), facilitating the entry of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) into respiratory cells [107], indicating the complex role of Nrf2 in SARS-CoV-2 infection.

The evidence highlights the complexity of using itaconate in the treatment of SARS-CoV-2. As an Nrf2 agonist, itaconate may increase cellular susceptibility to lipid toxicity induced by the spike protein. However, itaconate also inhibits SARS-CoV-2 replication through the Nrf2/HMOX1 pathway. Recently, it has been reported that 4-OI can block inflammation-associated coagulation by inhibiting type I interferon signaling and the release of tissue factor [108]. This finding offers a novel perspective on the therapeutic role of itaconate in managing inflammation.

Pulmonary fibrosis

Pulmonary fibrosis is a severe disease of unknown etiology characterized by excessive deposition of extracellular matrix and destruction of lung structure, leading to impaired gas exchange [109]. Studies have highlighted the role of alveolar macrophages (AMs) in regulating the pathogenic mechanisms underlying idiopathic pulmonary fibrosis (IPF), including lung defense, repair, surfactant handling, and inflammatory responses [110]. Given that itaconate is a major physiological regulator of global metabolic rewiring and the effector function of inflammatory macrophages, it holds potential therapeutic promise for influencing the development and progression of pulmonary fibrosis.

The expression of ACOD1 in alveolar macrophages (AMs) and itaconate levels in bronchoalveolar lavage fluid were both reduced in patients with idiopathic pulmonary fibrosis (IPF), indicating a significant alteration in the ACOD1/itaconate axis in fibrotic lung tissue [111]. In a bleomycin-induced mouse model of pulmonary fibrosis, specific knockout of ACOD1 resulted in exacerbation of the disease. Notably, adoptive transfer of wild-type monocyte-derived AMs into ACOD1-deficient mice restored the disease phenotype induced by bleomycin exposure. Additionally, the antifibrotic effects of inhaled itaconate were demonstrated in mice, confirming that itaconate is crucial for controlling the severity of pulmonary fibrosis [111]. Itaconate has also been shown to activate the Nrf2 pathway, which reduces the expression of thioredoxin interacting protein in the lung interstitium and mitigates fibroblast–myofibroblast differentiation (FMD)—a key cell phenotype in the development and progression of pulmonary fibrosis. Nrf2 inhibits transforming growth factor-β1-induced increases in FMD and ROS, thereby alleviating fibrosis [112]. It can be reasonably inferred that itaconate exerts beneficial effects on pulmonary fibrosis by activating the Nrf2 pathway. The potential mechanisms through which the Nrf2 pathway affects fibrotic disease have been well summarized [113]. As a natural metabolite and Nrf2 agonist, itaconate inhibits the initiation and progression of fibrosis by activating multiple downstream protective proteins via Nrf2, suggesting that itaconate could serve as a novel endogenous antifibrotic agent in the treatment of IPF.

Chronic obstructive pulmonary disease and asthma

Chronic obstructive pulmonary disease (COPD) is a condition characterized by persistent airflow limitation and is associated with a heightened chronic inflammatory response to irritants, while asthma is primarily associated with lower airway inflammation [1, 114]. Effective management of airway inflammation has been shown to alleviate symptoms in both COPD and asthma. The Nrf2 activator CPUY192018 can inhibit glycolysis and enhance antioxidant responses, thereby ameliorating inflammatory responses. CPUY192018 activates the Nrf2 pathway by disrupting the interaction between Keap1 and Nrf2, thereby reprogramming macrophage metabolism in an Nrf2-dependent manner [115]. Given that immune cells, such as macrophages, play a significant role in COPD pathogenesis [116], itaconate and its derivatives may also offer protective effects as effective Nrf2 activators. Further research is needed to determine whether itaconate administration can alleviate COPD. In asthma, loss of IRG1/itaconate was found to impair mitochondrial function in airway dendritic cells (DCs), leading to enhanced antigen priming. Administration of 4-OI [10 mg/kg intaperitoneally (i.p.)] restored mitochondrial redox balance in DCs and reduced the inflammatory response. Exogenous itaconate administration has been shown to decrease airway inflammation [64]. The inflammatory response is crucial in the pathogenesis of both COPD and asthma, and the anti-inflammatory and antioxidant effects of itaconate may offer novel therapeutic strategies.

Lung cancer

Lung cancer is a malignant tumor originating from the bronchial mucosa or glands of the lungs, characterized by rapid growth and high morbidity and mortality rates. Over the past 50 years, the incidence and mortality of lung cancer have significantly increased in many countries [117]. The etiology of lung cancer is still not fully understood. Lung cancer induces changes in the tumor microenvironment, including abnormal inflammatory cytokine profiles and oxidative stress. As an effective anti-inflammatory and antioxidant agent, itaconate may offer potential protective effects against the occurrence and progression of lung cancer.

The evidence linking itaconate to cancer is limited, and its role varies across different types of cancer. In some studies, itaconate has demonstrated anticancer effects. For instance, in colorectal cancer, which shares risk factors with lung cancer such as an unhealthy diet, smoking, and chronic inflammation, 4-OI inhibits aerobic glycolysis by targeting GAPDH, thereby promoting cuproptosis [118]. Additionally, DI has been shown to reduce the inflammatory state associated with ulcerative colitis and lower the risk of colitis-associated cancer [53]. In hepatocellular carcinoma, itaconate has exhibited anticancer effects comparable with those of 5-fluorouracil [119]. DI also has demonstrated anti-inflammatory and antitumor effects in human thymic cancer cell lines [120]. However, itaconate may also contribute to cancer progression. In non-small cell lung cancer, Nestin interacts with the Kelch domain of Keap1, leading to the escape of Nrf2 from Keap1-mediated degradation and subsequent promotion of antioxidant enzymes. The modulation of serotonin is emerging as a potential anticancer therapy, given that aberrant cancer antioxidant capacity may drive tumor malignancy. This suggests that the Nestin–Keap1–Nrf2 axis functions as a regulator of cellular redox homeostasis and confers resistance to oxidative stress in non-small cell lung cancer [121]. Itaconate has also been reported to downregulate PPARγ, a tumor suppressor, while upregulating anti-inflammatory cytokines in M2-like macrophages [122]. Furthermore, itaconate may promote tumor growth by suppressing cytotoxic CD8+ T cells and enhancing neutrophil resistance to ferroptosis [123, 124]. Itaconate’s role as an immune metabolic component in the crosstalk between tumor-associated macrophages and the tumor microenvironment indicates that dysregulated IRG1 expression may promote tumorigenesis by modulating antitumor immunity [29, 123]. Given the unclear relationship between itaconate and cancer, itaconate may not be a direct therapeutic target for lung cancer. Nonetheless, Wang et al. introduced novel CAR macrophages with enhanced polarization and antitumor functions through ACOD1 depletion [125], suggesting that itaconate might still have potential applications in lung cancer treatment in other contexts.

The therapeutic potential of itaconate derivatives

Itaconate is a pentacarbon dicarboxylic acid characterized by α,β-unsaturated olefins with mild electrophilic properties [32]. Although itaconate is generally considered a negatively charged polar metabolite with limited cell permeability, it can accumulate intracellularly in macrophages [48]. To overcome the permeability issue, various esterified derivatives of itaconate have been synthesized. Notably, 4-OI and DI are two commonly used derivatives with enhanced membrane permeability. Evidence shows that 4-OI can be converted intracellularly to itaconate by esterase, whereas DI does not methylate to itaconate [60, 126]. Other metabolites, such as fumarate, malonate, succinate, and phosphoenolpyruvate, share structural similarities with itaconate, which contributes to its role as a competitive inhibitor against these similar metabolites. Additionally, isomers of itaconate, including mesaconate and citraconate, exhibit similar immunomodulatory effects [59, 127]. In Table 2, we summarize the main characteristics of the most common itaconate derivatives, including their cell uptake, electrophilicity, and ability to convert to intracellular itaconate [48]. The immunomodulatory functions of itaconate and its derivatives involve several mechanisms, including SDH inhibition, Nrf2 activation, NLRP3 inflammasome inhibition, and glycolysis inhibition [128]. Recently, a novel post-translational modification known as lysine itaconylation has been reported [129], offering new perspectives for further functional studies of itaconate.

Table 2 Characteristics of different itaconate derivatives

So far, many itaconate derivatives have been tested. 4-OI, a cell-permeable itaconate derivative with an octyl ester tail, has shown potential protective effects in oxidative stress-related diseases [130, 131]. While 4-OI exhibits similar thiol reactivity to itaconate and can potentially be hydrolyzed to itaconate, it does not lead to intracellular accumulation of itaconate in LPS-activated IRG1-deficient macrophages [48]. This suggests that 4-OI may not effectively convert to intracellular itaconate, as itaconate can only arise from the de-esterification of derivatives in these macrophages. Nevertheless, 4-OI remains a functional Nrf2 activator [16]. In contrast, DI enhances LPS-mediated de novo synthesis of itaconate but does not convert to itaconate itself [126]. DI has increased electrophilicity compared with itaconate due to the esterification of the carboxyl group [132], and it is an effective and durable Nrf2 activator. The Nrf2-itaconate axis and antioxidant responses are crucial for pulmonary immune homeostasis. Another therapeutic potential of itaconate is its antiviral property, which can limit viral replication [79]. 4-Ethyl itaconate (4-EI), with a structure similar to DI but lower electrophilicity and higher polarity, cannot inhibit IκBζ. Differences in electrophilicity and other chemical properties among these derivatives can result in variations from the in vivo effects of itaconate. These differences highlight the need for further refinement in pharmaceutical processes to enhance the clinical translation of experimental results. In many lung diseases, itaconate demonstrates anti-inflammatory and anti-oxidative stress functions, which may offer significant implications for the treatment of infectious and inflammatory diseases. In Fig. 3, we summarized the therapeutic potential of itaconate and its derivatives in different lung diseases.

Fig. 3
figure 3

Itaconate and its derivatives can affect different lung diseases via various pathways, such as lung injury, asthma, fibrosis, lung cancer, etc. 4-OI and DI are the most commonly used itaconate derivatives in pulmonary diseases. Most studies focused on macrophages. 4-OI and DI attenuate acute lung injury via several pathways, such as TET2 inhibition, PI3K/AKT/NF-κB signaling inhibition, Nrf2-dependent ferroptosis inhiation or the NLRP3 inflammasome activation. Itaconate may reduce reactive oxygen species production, attenuate inflammatory responses and enhance the antimicrobial innate immunity through alkylation modification and inhibition of succinate dehydrogenase activity. Itaconate can also inhibit viral replication and bacterial growth. The tumor suppressor factor PPARγ can be suppressed by itaconate in tumor-associated macrophages. In addition, itaconate and its derivatives can function in other pulmonary cellular components, including fibroblasts, epithelial cells and dendritic cells by similar effects in macrophages, thus being involved in the pathogenesis of pulmonary fibrosis, lung cancer, and asthma. Black pointed arrows indicate promotion. Red blunt head arrows indicate inhibition

Many efforts have been made to enhance the delivery of itaconate. For instance, itaconate dissolved in saline has been infused before cerebral ischemia reperfusion in mouse models, demonstrating protective effects [70]. Intraperitoneal injection is the most commonly used delivery method in previous studies. Additionally, oropharyngeal administration has emerged as a potentially effective method, particularly for pulmonary diseases. A previous study reported that an inhaled dose of 0.25 mg/kg of itaconate, dissolved in phosphate-buffered saline, can ameliorate bleomycin-induced pulmonary fibrosis in mice [111]. Chen et al. developed a novel mitochondria-targeted supramolecular drug delivery system, 4-OI@Zn-NH-pyr, designed to scavenge both exogenous and endogenous ROS in joint inflammation [133]. Although the efficacy and efficiency of inhalation therapy with itaconate are supported by limited studies and require further validation, oropharyngeal inhalation of itaconate and its derivatives remains a promising and attractive approach. However, several issues need to be addressed before itaconate can be widely applied for lung diseases. While itaconate and its derivatives are functional immunometabolism regulators, they exhibit different modes of action. Both intraperitoneal and oropharyngeal administration methods have shown potential, but further refinement is needed. One concern is that pathogens might exploit the host immune response to sustain infection, which must be considered [134]. For systemic delivery methods, such as intravenous or intraperitoneal injection, itaconate might not accumulate in the lungs as intended, necessitating a targeted delivery vehicle. Inhaled drugs offer lower doses and reduced systemic side effects, but the formulation, dose, and delivery vehicle for itaconate still require further investigation. Despite these challenges, itaconate and its derivatives are promising therapeutic targets for lung diseases. They have been effectively tested in various lung disease models, as summarized in Table 3. Among these studies, LPS was the most common stimulus in animal models [24, 26, 52, 72, 73, 108, 135, 136], while viruses and other pathogens were used in several studies [64, 90, 104, 108, 137,138,139]. These models primarily assessed therapeutic effects in acute lung infection scenarios, with some research focusing on air pollution [140], chronic lung diseases [112, 141], or ventilator-induced lung injury [131].

Table 3 Recent studies involved itaconate and pulmonary disease

Future perspective and conclusion

Itaconate and its derivatives have emerged as significant regulators of inflammation and immune metabolism reprogramming. Typically, IRG1, induced by external stimuli like pathogens, leads to itaconate accumulation. This increased itaconate concentration triggers several changes in both immune and nonimmune cells, including SDH inhibition, Nrf2 activation, and electrophilic stress responses.

Given its potential as an anti-inflammatory therapeutic agent, significant efforts have been directed toward exploring itaconate and its derivatives in respiratory diseases, particularly in acute and chronic inflammatory lung conditions such as acute lung injury from sepsis and pulmonary fibrosis. While the efficacy of itaconate and its derivatives has been promising in various animal models, several issues remain unresolved before these compounds can be translated into clinical practice. One critical issue is determining the appropriate doses and concentrations of itaconate and its derivatives for clinical use. Animal studies have employed a wide range of doses, from 12.5 to 200 mg/kg. There is a concern that low doses may offer limited anti-inflammatory effects, while high doses could potentially lead to excessive inhibition of inflammation. Establishing optimal dosing parameters for different lung diseases requires further investigation. Another important consideration is the route of administration. Most studies have used intraperitoneal injection, but intravenous administration might be more suitable for clinical treatment due to its rapid onset and avoidance of first-pass metabolism by the liver. Inhalation is also a viable route, offering the potential for equivalent therapeutic effects at lower doses and bypassing hepatic metabolism. The dosing and formulation of itaconate and its derivatives may vary with different administration routes. Targeted delivery methods, such as modified liposomes, could further enhance the specificity and efficacy of itaconate therapy while minimizing side effects.

In conclusion, itaconate, as a metabolite derived from the TCA cycle, regulates inflammatory responses through various mechanisms and exhibits anti-inflammatory effects with low toxicity. It represents a promising therapeutic target for many lung diseases, particularly those related to inflammation. However, most evidence is derived from animal models, and further research, including clinical studies, is needed to establish the optimal doses, concentrations, and administration routes for clinical application.

Availability of data and materials

Not applicable.

Abbreviations

4-OI:

4-Octyl itaconate

ACOD1:

Aconitate decarboxylase 1

ARDS:

Acute respiratory distress syndrome

α-KG:

α-Ketoglutaric acid

CAD:

cis-aconitate decarboxylase

COPD:

Chronic obstructive pulmonary disease

CPT1A:

Carnitine palmitoyltransferase 1A

DCs:

Dendritic cells

DI:

Dimethyl itaconate

FMD:

Fibroblast–myofibroblast differentiation

GAPDH:

Glyceraldehyde 3 phosphate dehydrogenase

IFN-β:

Interferon-β

IRG1:

Immune response gene 1

IL-1β:

Interleukin-1β

JAK1:

Janus kinase 1

Keap1:

Kelch-like ECH-associated protein

LPS:

Lipopolysaccharide

NADPH:

Nicotinamide adenine dinucleotide phosphate

NETs:

Neutrophil extracellular traps

NF-κB:

Nuclear factor kappa-B

NLRP3:

NOD-like receptor protein 3

NO:

Nitric oxide

Nrf2:

Nuclear factor erythroid 2-related factor 2

IPF:

Idiopathic pulmonary fibrosis

PPARγ:

Peroxisome proliferator-activated receptor gamma

ROS:

Reactive oxygen species

SARS-CoV-2:

Severe acute respiratory syndrome coronavirus 2

SDH:

Succinate dehydrogenase

TCA:

Tricarboxylic acid

TFEB:

Transcription factor EB

References

  1. Lareau SC, Fahy B, Meek P, Wang A. Chronic Obstructive Pulmonary Disease (COPD). Am J Resp Crit Care. 2019;199:P1–2. https://doi.org/10.1164/rccm.1991P1.

    Article  Google Scholar 

  2. Mokra D. Acute lung injury - from pathophysiology to treatment. Physiol Res. 2020;69:S353–66. https://doi.org/10.33549/physiolres.934602.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Myllarniemi M, Kaarteenaho R. Pharmacological treatment of idiopathic pulmonary fibrosis - preclinical and clinical studies of pirfenidone, nintedanib, and N-acetylcysteine. Eur Cli Respir J. 2015. https://doi.org/10.3402/ecrj.v2.26385.

    Article  Google Scholar 

  4. Dong Q, Chen ES, Zhao C, Jin C. Host-microbiome interaction in lung cancer. Front Immunol. 2021;12:679829. https://doi.org/10.3389/fimmu.2021.679829.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Gan PXL, Liao W, Lim HF, Wong WSF. Dexamethasone protects against Aspergillus fumigatus-induced severe asthma via modulating pulmonary immunometabolism. Pharmacol Res. 2023;196:106929. https://doi.org/10.1016/j.phrs.2023.106929.

    Article  CAS  PubMed  Google Scholar 

  6. Ning L, Shishi Z, Bo W, Huiqing L. Targeting immunometabolism against acute lung injury. Clin Immunol. 2023;249:109289. https://doi.org/10.1016/j.clim.2023.109289.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Baup S. Ueber eine neue Pyrogen-Citronensäure, und über Benennung der Pyrogen-Säuren überhaupt. Annalen der Pharmacie. 2006;19:29–38. https://doi.org/10.1002/jlac.18360190107.

    Article  Google Scholar 

  8. Holmes FL. Hans Krebs: Architect of Intermediary Metabolism 1933–1937, Oxford University Press, 1993. https://doi.org/10.1093/oso/9780195076578.001.0001

    Book  Google Scholar 

  9. 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. https://doi.org/10.1016/j.nut.2004.08.007.

    Article  CAS  PubMed  Google Scholar 

  10. Strelko CL, Lu W, Dufort FJ, Seyfried TN, Chiles TC, Rabinowitz JD, Roberts MF. Itaconic acid is a mammalian metabolite induced during macrophage activation. J Am Chem Soc. 2011;133:16386–9. https://doi.org/10.1021/ja2070889.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Sugimoto M, Sakagami H, Yokote Y, Onuma H, Kaneko M, Mori M, Sakaguchi Y, Soga T, Tomita M. Non-targeted metabolite profiling in activated macrophage secretion. Metabolomics. 2011;8:624–33. https://doi.org/10.1007/s11306-011-0353-9.

    Article  CAS  Google Scholar 

  12. Michelucci A, Cordes T, Ghelfi J, Pailot A, Reiling N, Goldmann O, Binz T, Wegner A, Tallam A, Rausell A, Buttini M, Linster CL, Medina E, Balling R, Hiller K. Immune-responsive gene 1 protein links metabolism to immunity by catalyzing itaconic acid production. P Natl Acad Sci USA. 2013;110:7820–5. https://doi.org/10.1073/pnas.1218599110.

    Article  Google Scholar 

  13. Basler T, Jeckstadt S, Valentin-Weigand P, Goethe R. Mycobacterium paratuberculosis, Mycobacterium smegmatis, and lipopolysaccharide induce different transcriptional and post-transcriptional regulation of the IRG1 gene in murine macrophages. J Leukocyte Biol. 2006;79:628–38. https://doi.org/10.1189/jlb.0905520.

    Article  CAS  PubMed  Google Scholar 

  14. Thomas DM, Francescutti-Verbeem DM, Kuhn DM. Gene expression profile of activated microglia under conditions associated with dopamine neuronal damage. Faseb J. 2006;20:515–7. https://doi.org/10.1096/fj.05-4873fje.

    Article  CAS  PubMed  Google Scholar 

  15. Zaslona Z, ONeill LAJ. Cytokine-like roles for metabolites in immunity. Molecular cell. 2020;78:814–23. https://doi.org/10.1016/j.molcel.2020.04.002.

    Article  CAS  PubMed  Google Scholar 

  16. Mills EL, Ryan DG, Prag HA, Dikovskaya D, Menon D, Zaslona Z, Jedrychowski MP, Costa ASH, Higgins M, Hams E, Szpyt J, Runtsch MC, King MS, McGouran JF, Fischer R, Kessler BM, McGettrick AF, Hughes MM, Carroll RG, Booty LM, Knatko EV, Meakin PJ, Ashford MLJ, Modis LK, Brunori G, Sevin DC, Fallon PG, Caldwell ST, Kunji ERS, Chouchani ET, Frezza C, Dinkova-Kostova AT, Hartley RC, Murphy MP, ONeill LA. Itaconate is an anti-inflammatory metabolite that activates Nrf2 via alkylation of KEAP1. Nature. 2018;556:113–7. https://doi.org/10.1038/nature25986.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Cordes T, Wallace M, Michelucci A, Divakaruni AS, Sapcariu SC, Sousa C, Koseki H, Cabrales P, Murphy AN, Hiller K, Metallo CM. Immunoresponsive gene 1 and itaconate inhibit succinate dehydrogenase to modulate intracellular succinate levels. J Biol Chem. 2016;291:14274–84. https://doi.org/10.1074/jbc.M115.685792.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Lampropoulou V, Sergushichev A, Bambouskova M, Nair S, Vincent EE, Loginicheva E, Cervantes-Barragan L, Ma X, Huang SC, Griss T, Weinheimer CJ, Khader S, Randolph GJ, Pearce EJ, Jones RG, Diwan A, Diamond MS, Artyomov MN. Itaconate links inhibition of succinate dehydrogenase with macrophage metabolic remodeling and regulation of inflammation. Cell Metab. 2016;24:158–66. https://doi.org/10.1016/j.cmet.2016.06.004.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Nemeth B, Doczi J, Csete D, Kacso G, Ravasz D, Adams D, Kiss G, Nagy AM, Horvath G, Tretter L, Mocsai A, Csepanyi-Komi R, Iordanov I, Adam-Vizi V, Chinopoulos C. Abolition of mitochondrial substrate-level phosphorylation by itaconic acid produced by LPS-induced Irg1 expression in cells of murine macrophage lineage. Faseb J. 2016;30:286–300. https://doi.org/10.1096/fj.15-279398.

    Article  CAS  PubMed  Google Scholar 

  20. Tang C, Wang X, Xie Y, Cai X, Yu N, Hu Y, Zheng Z. 4-Octyl itaconate activates Nrf2 signaling to inhibit pro-inflammatory cytokine production in peripheral blood mononuclear cells of systemic lupus erythematosus patients. Cell Physiol Biochem. 2018;51:979–90. https://doi.org/10.1159/000495400.

    Article  CAS  PubMed  Google Scholar 

  21. Bambouskova M, Gorvel L, Lampropoulou V, Sergushichev A, Loginicheva E, Johnson K, Korenfeld D, Mathyer ME, Kim H, Huang LH, Duncan D, Bregman H, Keskin A, Santeford A, Apte RS, Sehgal R, Johnson B, Amarasinghe GK, Soares MP, Satoh T, Akira S, Hai T, de Guzman Strong C, Auclair K, Roddy TP, Biller SA, Jovanovic M, Klechevsky E, Stewart KM, Randolph GJ, Artyomov MN. Electrophilic properties of itaconate and derivatives regulate the IkappaBzeta-ATF3 inflammatory axis. Nature. 2018;556:501–4. https://doi.org/10.1038/s41586-018-0052-z.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Hooftman A, Angiari S, Hester S, Corcoran SE, Runtsch MC, Ling C, Ruzek MC, Slivka PF, McGettrick AF, Banahan K, Hughes MM, Irvine AD, Fischer R, O’Neill LAJ. The immunomodulatory metabolite itaconate modifies NLRP3 and inhibits inflammasome activation. Cell Metab. 2020;32(468–478):e467. https://doi.org/10.1016/j.cmet.2020.07.016.

    Article  CAS  Google Scholar 

  23. Qin W, Qin K, Zhang Y, Jia W, Chen Y, Cheng B, Peng L, Chen N, Liu Y, Zhou W, Wang YL, Chen X, Wang C. S-glycosylation-based cysteine profiling reveals regulation of glycolysis by itaconate. Nat Chem Biol. 2019;15:983–91. https://doi.org/10.1038/s41589-019-0323-5.

    Article  CAS  PubMed  Google Scholar 

  24. Chen LL, Morcelle C, Cheng ZL, Chen X, Xu Y, Gao Y, Song J, Li Z, Smith MD, Shi M, Zhu Y, Zhou N, Cheng M, He C, Liu KY, Lu G, Zhang L, Zhang C, Zhang J, Sun Y, Qi T, Lyu Y, Ren ZZ, Tan XM, Yin J, Lan F, Liu Y, Yang H, Qian M, Duan C, Chang X, Zhou Y, Shen L, Baldwin AS, Guan KL, Xiong Y, Ye D. Itaconate inhibits TET DNA dioxygenases to dampen inflammatory responses. Nat Cell Biol. 2022;24:353–63. https://doi.org/10.1038/s41556-022-00853-8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Runtsch MC, Angiari S, Hooftman A, Wadhwa R, Zhang Y, Zheng Y, Spina JS, Ruzek MC, Argiriadi MA, McGettrick AF, Mendez RS, Zotta A, Peace CG, Walsh A, Chirillo R, Hams E, Fallon PG, Jayamaran R, Dua K, Brown AC, Kim RY, Horvat JC, Hansbro PM, Wang C, O’Neill LAJ. Itaconate and itaconate derivatives target JAK1 to suppress alternative activation of macrophages. Cell Metab. 2022;34(487–501):e488. https://doi.org/10.1016/j.cmet.2022.02.002.

    Article  CAS  Google Scholar 

  26. Zhang Z, Chen C, Yang F, Zeng YX, Sun P, Liu P, Li X. Itaconate is a lysosomal inducer that promotes antibacterial innate immunity. Mol cell. 2022;82(2844–2857):e2810. https://doi.org/10.1016/j.molcel.2022.05.009.

    Article  CAS  Google Scholar 

  27. Peace CG, O’Neill LA. The role of itaconate in host defense and inflammation. J Clin Invest. 2022;132:e148548. https://doi.org/10.1172/JCI148548.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Winterhoff M, Chen F, Sahini N, Ebensen T, Kuhn M, Kaever V, Bahre H, Pessler F. Establishment, validation, and initial application of a sensitive LC-MS/MS assay for quantification of the naturally occurring isomers itaconate, mesaconate, and citraconate. Metabolites. 2021;11:270. https://doi.org/10.3390/metabo11050270.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Wu R, Chen F, Wang N, Tang D, Kang R. ACOD1 in immunometabolism and disease. Cell Mol Immunol. 2020;17:822–33. https://doi.org/10.1038/s41423-020-0489-5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Xiao W, Wang L, Xiao R, Wu M, Tan J, He Y. Expression profile of human immune-responsive gene 1 and generation and characterization of polyclonal antiserum. Mol Cell Biochem. 2011;353:177–87. https://doi.org/10.1007/s11010-011-0784-7.

    Article  CAS  PubMed  Google Scholar 

  31. Orlowska K, Fling RR, Nault R, Sink WJ, Schilmiller AL, Zacharewski T. Dioxin-elicited decrease in cobalamin redirects propionyl-CoA metabolism to the beta-oxidation-like pathway resulting in acrylyl-CoA conjugate buildup. J Biol Chem. 2022;298:102301. https://doi.org/10.1016/j.jbc.2022.102301.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. O’Neill LAJ, Artyomov MN. Itaconate: the poster child of metabolic reprogramming in macrophage function. Nat Rev Immunol. 2019;19:273–81. https://doi.org/10.1038/s41577-019-0128-5.

    Article  CAS  PubMed  Google Scholar 

  33. Tallam A, Perumal TM, Antony PM, Jager C, Fritz JV, Vallar L, Balling R, Del Sol A, Michelucci A. Gene regulatory network inference of immunoresponsive gene 1 (IRG1) identifies interferon regulatory factor 1 (IRF1) as its transcriptional regulator in mammalian macrophages. Plos One. 2016;11:e0149050. https://doi.org/10.1371/journal.pone.0149050.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Schuster EM, Epple MW, Glaser KM, Mihlan M, Lucht K, Zimmermann JA, Bremser A, Polyzou A, Obier N, Cabezas-Wallscheid N, Trompouki E, Ballabio A, Vogel J, Buescher JM, Westermann AJ, Rambold AS. TFEB induces mitochondrial itaconate synthesis to suppress bacterial growth in macrophages. Nat Metab. 2022;4:856–66. https://doi.org/10.1038/s42255-022-00605-w.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Chen F, Wu R, Liu J, Kang R, Li J, Tang D. The STING1-MYD88 complex drives ACOD1/IRG1 expression and function in lethal innate immunity. Iscience. 2022;25:104561. https://doi.org/10.1016/j.isci.2022.104561.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Bailey JD, Diotallevi M, Nicol T, McNeill E, Shaw A, Chuaiphichai S, Hale A, Starr A, Nandi M, Stylianou E, McShane H, Davis S, Fischer R, Kessler BM, McCullagh J, Channon KM, Crabtree MJ. Nitric oxide modulates metabolic remodeling in inflammatory macrophages through TCA cycle regulation and itaconate accumulation. Cell Rep. 2019;28:218–30. https://doi.org/10.1016/j.celrep.2019.06.018.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Chen YJ, Li GN, Li XJ, Wei LX, Fu MJ, Cheng ZL, Yang Z, Zhu GQ, Wang XD, Zhang C, Zhang JY, Sun YP, Saiyin H, Zhang J, Liu WR, Zhu WW, Guan KL, Xiong Y, Yang Y, Ye D, Chen LL. Targeting IRG1 reverses the immunosuppressive function of tumor-associated macrophages and enhances cancer immunotherapy. Sci Adv. 2023;9:eadg0654. https://doi.org/10.1126/sciadv.adg0654.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Van Quickelberghe E, Martens A, Goeminne LJE, Clement L, van Loo G, Gevaert K. Identification of immune-responsive gene 1 (IRG1) as a target of A20. J Proteome Res. 2018;17:2182–91. https://doi.org/10.1021/acs.jproteome.8b00139.

    Article  CAS  PubMed  Google Scholar 

  39. Ganta VC, Choi MH, Kutateladze A, Fox TE, Farber CR, Annex BH. A MicroRNA93-interferon regulatory factor-9-immunoresponsive gene-1-itaconic acid pathway modulates M2-like macrophage polarization to revascularize ischemic muscle. Circulation. 2017;135:2403–25. https://doi.org/10.1161/CIRCULATIONAHA.116.025490.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Scheurlen KM, Chariker JH, Kanaan Z, Littlefield AB, George JB, Seraphine C, Rochet A, Rouchka EC, Galandiuk S. The NOTCH4-GATA4-IRG1 axis as a novel target in early-onset colorectal cancer. Cytokine Growth F R. 2022;67:25–34. https://doi.org/10.1016/j.cytogfr.2022.06.002.

    Article  CAS  Google Scholar 

  41. Huang J, Jiang Y, Ji R, Jia Y, Wang S, Zhou Z, Wang S, Wang J, Yang Q, Bai H, Zhu X, Jiang B, Ben J, Zhang H, Li X, Chen Q. Macrophage scavenger receptor A1 antagonizes abdominal aortic aneurysm via upregulating IRG1. Biochem Pharmacol. 2023;213:115631. https://doi.org/10.1016/j.bcp.2023.115631.

    Article  CAS  PubMed  Google Scholar 

  42. Nelson VL, Nguyen HCB, Garcia-Canaveras JC, Briggs ER, Ho WY, DiSpirito JR, Marinis JM, Hill DA, Lazar MA. PPARgamma is a nexus controlling alternative activation of macrophages via glutamine metabolism. Genes Dev. 2018;32:1035–44. https://doi.org/10.1101/gad.312355.118.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. De Souza DP, Achuthan A, Lee MK, Binger KJ, Lee MC, Davidson S, Tull DL, McConville MJ, Cook AD, Murphy AJ, Hamilton JA, Fleetwood AJ. Autocrine IFN-I inhibits isocitrate dehydrogenase in the TCA cycle of LPS-stimulated macrophages. J Clin Invest. 2019;129:4239–44. https://doi.org/10.1172/JCI127597.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Yang W, Wang Y, Zhang P, Sun X, Chen X, Yu J, Shi L, Yin Y, Tao K, Li R. Immune-responsive gene 1 protects against liver injury caused by concanavalin A via the activation Nrf2/HO-1 pathway and inhibition of ROS activation pathways. Free Radic Bio Med. 2022;182:108–18. https://doi.org/10.1016/j.freeradbiomed.2022.02.030.

    Article  CAS  Google Scholar 

  45. Pan J, Zhao X, Lin C, Xu H, Yin Z, Liu T, Zhang S. Immune responsive gene 1, a novel oncogene, increases the growth and tumorigenicity of glioma. Oncol Rep. 2014;32:1957–66. https://doi.org/10.3892/or.2014.3474.

    Article  CAS  PubMed  Google Scholar 

  46. Yang W, Wang Y, Wang T, Li C, Shi L, Zhang P, Yin Y, Tao K, Li R. Protective effects of IRG1/itaconate on acute colitis through the inhibition of gasdermins-mediated pyroptosis and inflammation response. Genes Dis. 2023;10:1552–63. https://doi.org/10.1016/j.gendis.2022.05.039.

    Article  CAS  PubMed  Google Scholar 

  47. Singh S, Singh PK, Jha A, Naik P, Joseph J, Giri S, Kumar A. Integrative metabolomics and transcriptomics identifies itaconate as an adjunct therapy to treat ocular bacterial infection. Cell Rep Med. 2021;2:100277. https://doi.org/10.1016/j.xcrm.2021.100277.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Swain A, Bambouskova M, Kim H, Andhey PS, Duncan D, Auclair K, Chubukov V, Simons DM, Roddy TP, Stewart KM, Artyomov MN. Comparative evaluation of itaconate and its derivatives reveals divergent inflammasome and type I interferon regulation in macrophages. Nat Metab. 2020;2:594–602. https://doi.org/10.1038/s42255-020-0210-0.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Cordes T, Metallo CM. Itaconate alters succinate and coenzyme a metabolism via inhibition of mitochondrial complex II and methylmalonyl-CoA mutase. Metabolites. 2021. https://doi.org/10.3390/metabo11020117.

    Article  PubMed  PubMed Central  Google Scholar 

  50. Mainali R, Buechler N, Otero C, Edwards L, Key CC, Furdui C, Quinn MA. Itaconate stabilizes CPT1a to enhance lipid utilization during inflammation. Elife. 2024. https://doi.org/10.7554/eLife.92420.

    Article  PubMed  PubMed Central  Google Scholar 

  51. Olagnier D, Brandtoft AM, Gunderstofte C, Villadsen NL, Krapp C, Thielke AL, Laustsen A, Peri S, Hansen AL, Bonefeld L, Thyrsted J, Bruun V, Iversen MB, Lin L, Artegoitia VM, Su C, Yang L, Lin R, Balachandran S, Luo Y, Nyegaard M, Marrero B, Goldbach-Mansky R, Motwani M, Ryan DG, Fitzgerald KA, O’Neill LA, Hollensen AK, Damgaard CK, de Paoli FV, Bertram HC, Jakobsen MR, Poulsen TB, Holm CK. Nrf2 negatively regulates STING indicating a link between antiviral sensing and metabolic reprogramming. Nat Commun. 2018;9:3506. https://doi.org/10.1038/s41467-018-05861-7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Xin Y, Zou L, Lang S. 4-Octyl itaconate (4-OI) attenuates lipopolysaccharide-induced acute lung injury by suppressing PI3K/Akt/NF-kappaB signaling pathways in mice. Exp Ther Med. 2021;21:141. https://doi.org/10.3892/etm.2020.9573.

    Article  CAS  PubMed  Google Scholar 

  53. Chen M, Sun H, Boot M, Shao L, Chang SJ, Wang W, Lam TT, Lara-Tejero M, Rego EH, Galan JE. Itaconate is an effector of a Rab GTPase cell-autonomous host defense pathway against Salmonella. Science. 2020;369:450–5. https://doi.org/10.1126/science.aaz1333.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Nair S, Huynh JP, Lampropoulou V, Loginicheva E, Esaulova E, Gounder AP, Boon ACM, Schwarzkopf EA, Bradstreet TR, Edelson BT, Artyomov MN, Stallings CL, Diamond MS. Irg1 expression in myeloid cells prevents immunopathology during M. tuberculosis infection. J Exp Med. 2018;215:1035–45. https://doi.org/10.1084/jem.20180118.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Liu X, Shi B, Suo R, Xiong S, Wang X, Liang X, Li X, Li G. Itaconate regulates macrophage function through stressful iron-sulfur cluster disrupting and iron metabolism rebalancing. Faseb J. 2021;35:e21936. https://doi.org/10.1096/fj.202100726RR.

    Article  CAS  PubMed  Google Scholar 

  56. Muri J, Wolleb H, Broz P, Carreira EM, Kopf M. Electrophilic Nrf2 activators and itaconate inhibit inflammation at low dose and promote IL-1beta production and inflammatory apoptosis at high dose. Redox Biol. 2020;36:101647. https://doi.org/10.1016/j.redox.2020.101647.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Daniels BP, Kofman SB, Smith JR, Norris GT, Snyder AG, Kolb JP, Gao X, Locasale JW, Martinez J, Gale M Jr, Loo YM, Oberst A. The nucleotide sensor ZBP1 AND Kinase RIPK3 induce the enzyme IRG1 to promote an antiviral metabolic state in neurons. Immunity. 2019;50:64–76. https://doi.org/10.1016/j.immuni.2018.11.017.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Azzimato V, Chen P, Barreby E, Morgantini C, Levi L, Vankova A, Jager J, Sulen A, Diotallevi M, Shen JX, Miller A, Ellis E, Ryden M, Naslund E, Thorell A, Lauschke VM, Channon KM, Crabtree MJ, Haschemi A, Craige SM, Mori M, Spallotta F, Aouadi M. Hepatic miR-144 drives fumarase activity preventing NRF2 activation during obesity. Gastroenterology. 2021;161:1982–97. https://doi.org/10.1053/j.gastro.2021.08.030.

    Article  CAS  PubMed  Google Scholar 

  59. He W, Henne A, Lauterbach M, Geissmar E, Nikolka F, Kho C, Heinz A, Dostert C, Grusdat M, Cordes T, Harm J, Goldmann O, Ewen A, Verschueren C, Blay-Cadanet J, Geffers R, Garritsen H, Kneiling M, Holm CK, Metallo CM, Medina E, Abdullah Z, Latz E, Brenner D, Hiller K. Mesaconate is synthesized from itaconate and exerts immunomodulatory effects in macrophages. Nat Metab. 2022;4:524–33. https://doi.org/10.1038/s42255-022-00565-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Liao ST, Han C, Xu DQ, Fu XW, Wang JS, Kong LY. 4-Octyl itaconate inhibits aerobic glycolysis by targeting GAPDH to exert anti-inflammatory effects. Nat Commun. 2019;10:5091. https://doi.org/10.1038/s41467-019-13078-5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Huang SS, Guo DY, Jia BB, Cai GL, Yan J, Lu Y, Yang ZX. Dimethyl itaconate alleviates the pyroptosis of macrophages through oxidative stress. Bmc Immunol. 2021;22:72. https://doi.org/10.1186/s12865-021-00463-3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Waqas SF, Sohail A, Nguyen AHH, Usman A, Ludwig T, Wegner A, Malik MNH, Schuchardt S, Geffers R, Winterhoff M, Merkert S, Martin U, Olmer R, Lachmann N, Pessler F. ISG15 deficiency features a complex cellular phenotype that responds to treatment with itaconate and derivatives. Clin Transl Med. 2022;12:e931. https://doi.org/10.1002/ctm2.931.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Weiss JM, Davies LC, Karwan M, Ileva L, Ozaki MK, Cheng RY, Ridnour LA, Annunziata CM, Wink DA, McVicar DW. Itaconic acid mediates crosstalk between macrophage metabolism and peritoneal tumors. J Clin Invest. 2018;128:3794–805. https://doi.org/10.1172/JCI99169.

    Article  PubMed  PubMed Central  Google Scholar 

  64. Jaiswal AK, Yadav J, Makhija S, Mazumder S, Mitra AK, Suryawanshi A, Sandey M, Mishra A. Irg1/itaconate metabolic pathway is a crucial determinant of dendritic cells immune-priming function and contributes to resolute allergen-induced airway inflammation. Mucosal Immunol. 2022;15:301–13. https://doi.org/10.1038/s41385-021-00462-y.

    Article  CAS  PubMed  Google Scholar 

  65. Li Y, Zhang P, Wang C, Han C, Meng J, Liu X, Xu S, Li N, Wang Q, Shi X, Cao X. Immune responsive gene 1 (IRG1) promotes endotoxin tolerance by increasing A20 expression in macrophages through reactive oxygen species. J Biol Chem. 2013;288:16225–34. https://doi.org/10.1074/jbc.M113.454538.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Liu X, Wu XP, Zhu XL, Li T, Liu Y. IRG1 increases MHC class I level in macrophages through STAT-TAP1 axis depending on NADPH oxidase mediated reactive oxygen species. Int Immunopharmacol. 2017;48:76–83. https://doi.org/10.1016/j.intimp.2017.04.012.

    Article  CAS  PubMed  Google Scholar 

  67. Zheng Y, Chen Z, She C, Lin Y, Hong Y, Shi L, Zhang Y, Cao P, Xu X. Four-octyl itaconate activates Nrf2 cascade to protect osteoblasts from hydrogen peroxide-induced oxidative injury. Cell Death Dis. 2020;11:772. https://doi.org/10.1038/s41419-020-02987-9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Liu H, Feng Y, Xu M, Yang J, Wang Z, Di G. Four-octyl itaconate activates Keap1-Nrf2 signaling to protect neuronal cells from hydrogen peroxide. Cell Commun Signal. 2018;16:81. https://doi.org/10.1186/s12964-018-0294-2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Yi Z, Deng M, Scott MJ, Fu G, Loughran PA, Lei Z, Li S, Sun P, Yang C, Li W, Xu H, Huang F, Billiar TR. Immune-responsive gene 1/itaconate activates nuclear factor erythroid 2-related factor 2 in hepatocytes to protect against liver ischemia-reperfusion injury. Hepatology. 2020;72:1394–411. https://doi.org/10.1002/hep.31147.

    Article  CAS  PubMed  Google Scholar 

  70. Cordes T, Lucas A, Divakaruni AS, Murphy AN, Cabrales P, Metallo CM. Itaconate modulates tricarboxylic acid and redox metabolism to mitigate reperfusion injury. Mol Metab. 2020;32:122–35. https://doi.org/10.1016/j.molmet.2019.11.019.

    Article  CAS  PubMed  Google Scholar 

  71. Seim GL, Britt EC, John SV, Yeo FJ, Johnson AR, Eisenstein RS, Pagliarini DJ, Fan J. Two-stage metabolic remodelling in macrophages in response to lipopolysaccharide and interferon-gamma stimulation. Nat Metab. 2019;1:731–42. https://doi.org/10.1038/s42255-019-0083-2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. He R, Liu B, Xiong R, Geng B, Meng H, Lin W, Hao B, Zhang L, Wang W, Jiang W, Li N, Geng Q. Itaconate inhibits ferroptosis of macrophage via Nrf2 pathways against sepsis-induced acute lung injury. Cell Death Discov. 2022;8:43. https://doi.org/10.1038/s41420-021-00807-3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Wu YT, Xu WT, Zheng L, Wang S, Wei J, Liu MY, Zhou HP, Li QF, Shi X, Lv X. 4-octyl itaconate ameliorates alveolar macrophage pyroptosis against ARDS via rescuing mitochondrial dysfunction and suppressing the cGAS/STING pathway. Int Immunopharmacol. 2023;118:110104. https://doi.org/10.1016/j.intimp.2023.110104.

    Article  CAS  PubMed  Google Scholar 

  74. Bambouskova M, Potuckova L, Paulenda T, Kerndl M, Mogilenko DA, Lizotte K, Swain A, Hayes S, Sheldon RD, Kim H, Kapadnis U, Ellis AE, Isaguirre C, Burdess S, Laha A, Amarasinghe GK, Chubukov V, Roddy TP, Diamond MS, Jones RG, Simons DM, Artyomov MN. Itaconate confers tolerance to late NLRP3 inflammasome activation. Cell Rep. 2021;34:108756. https://doi.org/10.1016/j.celrep.2021.108756.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Liu K, Huang J, Liu J, Klionsky DJ, Kang R, Tang D. Induction of autophagy-dependent ferroptosis to eliminate drug-tolerant human retinoblastoma cells. Cell Death Dis. 2022;13:521. https://doi.org/10.1038/s41419-022-04974-8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Mehra MR, Crespo-Leiro MG, Dipchand A, Ensminger SM, Hiemann NE, Kobashigawa JA, Madsen J, Parameshwar J, Starling RC, Uber PA. International Society for Heart and Lung Transplantation working formulation of a standardized nomenclature for cardiac allograft vasculopathy-2010. J Heart Lung Transpl. 2010;29:717–27. https://doi.org/10.1016/j.healun.2010.05.017.

    Article  Google Scholar 

  77. Ortiz-Munoz G, Mallavia B, Bins A, Headley M, Krummel MF, Looney MR. Aspirin-triggered 15-epi-lipoxin A4 regulates neutrophil-platelet aggregation and attenuates acute lung injury in mice. Blood. 2014;124:2625–34. https://doi.org/10.1182/blood-2014-03-562876.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Tasaka S, Amaya F, Hashimoto S, Ishizaka A. Roles of oxidants and redox signaling in the pathogenesis of acute respiratory distress syndrome. Antioxid Redox Sign. 2008;10:739–53. https://doi.org/10.1089/ars.2007.1940.

    Article  CAS  Google Scholar 

  79. Michalaki C, Albers GJ, Byrne AJ. Itaconate as a key regulator of respiratory disease. Clin Exp Immunol. 2024;215:120–5. https://doi.org/10.1093/cei/uxad127.

    Article  PubMed  Google Scholar 

  80. Xu L, Cai J, Li C, Yang M, Duan T, Zhao Q, Xi Y, Sun L, He L, Tang C, Sun L. 4-Octyl itaconate attenuates LPS-induced acute kidney injury by activating Nrf2 and inhibiting STAT3 signaling. Mol Med. 2023;29:58. https://doi.org/10.1186/s10020-023-00631-8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Ni L, Xiao J, Zhang D, Shao Z, Huang C, Wang S, Wu Y, Tian N, Sun L, Wu A, Zhou Y, Wang X, Zhang X. Immune-responsive gene 1/itaconate activates nuclear factor erythroid 2-related factor 2 in microglia to protect against spinal cord injury in mice. Cell Death Dis. 2022;13:140. https://doi.org/10.1038/s41419-022-04592-4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Liu G, Wu Y, Jin S, Sun J, Wan BB, Zhang J, Wang Y, Gao ZQ, Chen D, Li S, Pang Q, Wang Z. Itaconate ameliorates methicillin-resistant Staphylococcus aureus-induced acute lung injury through the Nrf2/ARE pathway. Ann Transl Med. 2021;9:712. https://doi.org/10.21037/atm-21-1448.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Darvish Khadem M, Tabandeh MR, Haschemi A, Kheirollah A, Shahriari A. Dimethyl itaconate reprograms neurotoxic to neuroprotective primary astrocytes through the regulation of NLRP3 inflammasome and NRF2/HO-1 pathways. Mol Cell Neurosci. 2022;122:103758. https://doi.org/10.1016/j.mcn.2022.103758.

    Article  CAS  PubMed  Google Scholar 

  84. Yang S, Zhang X, Zhang H, Lin X, Chen X, Zhang Y, Lin X, Huang L, Zhuge Q. Dimethyl itaconate inhibits LPS-induced microglia inflammation and inflammasome-mediated pyroptosis via inducing autophagy and regulating the Nrf-2/HO-1 signaling pathway. Mol Med Rep. 2021. https://doi.org/10.3892/mmr.2021.12311.

    Article  PubMed  PubMed Central  Google Scholar 

  85. Burczyk G, Cichon I, Kolaczkowska E. Itaconate suppresses formation of Neutrophil Extracellular Traps (NETs): involvement of hypoxia-inducible factor 1alpha (Hif-1alpha) and heme oxygenase (HO-1). Front Immunol. 2022;13:864638. https://doi.org/10.3389/fimmu.2022.864638.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Zhu L, Wu Z, Liu Y, Ming Y, Xie P, Jiang M, Qi Y. Acod1/itaconate activates Nrf2 in pulmonary microvascular endothelial cells to protect against the obesity-induced pulmonary microvascular endotheliopathy. Respir Res. 2024;25:205. https://doi.org/10.1186/s12931-024-02827-w.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Liu C, Fu C, Sun Y, You Y, Wang T, Zhang Y, Xia H, Wang X. Itaconic acid regulation of TFEB-mediated autophagy flux alleviates hyperoxia-induced bronchopulmonary dysplasia. Redox Biol. 2024;72:103115. https://doi.org/10.1016/j.redox.2024.103115.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Gu L, Lin J, Wang Q, Li C, Peng X, Fan Y, Lu C, Lin H, Niu Y, Zhu G, Zhao G. Dimethyl itaconate protects against fungal keratitis by activating the Nrf2/HO-1 signaling pathway. Immunol Cell Biol. 2020;98:229–41. https://doi.org/10.1111/imcb.12316.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Vassileva M, Malusa E, Eichler-Lobermann B, Vassilev N. Aspegillus terreus: from soil to industry and back. Microorganisms. 2020. https://doi.org/10.3390/microorganisms8111655.

    Article  PubMed  PubMed Central  Google Scholar 

  90. Sohail A, Iqbal AA, Sahini N, Chen F, Tantawy M, Waqas SFH, Winterhoff M, Ebensen T, Schultz K, Geffers R, Schughart K, Preusse M, Shehata M, Bahre H, Pils MC, Guzman CA, Mostafa A, Pleschka S, Falk C, Michelucci A, Pessler F. Itaconate and derivatives reduce interferon responses and inflammation in influenza A virus infection. PLoS Pathog. 2022;18:e1010219. https://doi.org/10.1371/journal.ppat.1010219.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Waqas FH, Shehata M, Elgaher WAM, Lacour A, Kurmasheva N, Begnini F, Kiib AE, Dahlmann J, Chen C, Pavlou A, Poulsen TB, Merkert S, Martin U, Olmer R, Olagnier D, Hirsch AKH, Pleschka S, Pessler F. NRF2 activators inhibit influenza A virus replication by interfering with nucleo-cytoplasmic export of viral RNPs in an NRF2-independent manner. PLoS Pathog. 2023;19:e1011506. https://doi.org/10.1371/journal.ppat.1011506.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Ribo-Molina P, Weiss HJ, Susma B, van Nieuwkoop S, Persoons L, Zheng Y, Ruzek M, Daelemans D, Fouchier RAM, O’Neill LAJ, van den Hoogen BG. 4-Octyl itaconate reduces influenza A replication by targeting the nuclear export protein CRM1. J Virol. 2023;97:e0132523. https://doi.org/10.1128/jvi.01325-23.

    Article  CAS  PubMed  Google Scholar 

  93. Zhu X, Guo Y, Liu Z, Yang J, Tang H, Wang Y. Itaconic acid exerts anti-inflammatory and antibacterial effects via promoting pentose phosphate pathway to produce ROS. Sci Rep. 2021;11:18173. https://doi.org/10.1038/s41598-021-97352-x.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Xie LY, Xu YB, Ding XQ, Liang S, Li DL, Fu AK, Zhan XA. Itaconic acid and dimethyl itaconate exert antibacterial activity in carbon-enriched environments through the TCA cycle. Biomed Pharmacother. 2023;167:115487. https://doi.org/10.1016/j.biopha.2023.115487.

    Article  CAS  PubMed  Google Scholar 

  95. Nguyen TV, Alfaro AC, Young T, Green S, Zarate E, Merien F. Itaconic acid inhibits growth of a pathogenic marine Vibrio strain: a metabolomics approach. Sci Rep. 2019;9:5937. https://doi.org/10.1038/s41598-019-42315-6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Zhang Y, Qin W, Liu D, Liu Y, Wang C. Chemoproteomic profiling of itaconations in Salmonella. Chem Sci. 2021;12:6059–63. https://doi.org/10.1039/d1sc00660f.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Bhusal RP, Bashiri G, Kwai BXC, Sperry J, Leung IKH. Targeting isocitrate lyase for the treatment of latent tuberculosis. Drug Discov Today. 2017;22:1008–16. https://doi.org/10.1016/j.drudis.2017.04.012.

    Article  CAS  PubMed  Google Scholar 

  98. Loi VV, Busche T, Kuropka B, Muller S, Methling K, Lalk M, Kalinowski J, Antelmann H. Staphylococcus aureus adapts to the immunometabolite itaconic acid by inducing acid and oxidative stress responses including S-bacillithiolations and S-itaconations. Free Radic Bio Med. 2023;208:859–76. https://doi.org/10.1016/j.freeradbiomed.2023.09.031.

    Article  CAS  Google Scholar 

  99. Ki N, Kim J, Jo I, Hyun Y, Ryu S, Ha NC. Isocitrate binds to the itaconic acid-responsive LysR-type transcriptional regulator RipR in Salmonella pathogenesis. J Biol Chem. 2022;298:102562. https://doi.org/10.1016/j.jbc.2022.102562.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Zhu N, Zhang D, Wang W, Li X, Yang B, Song J, Zhao X, Huang B, Shi W, Lu R, Niu P, Zhan F, Ma X, Wang D, Xu W, Wu G, Gao GF, Tan W. Novel coronavirus and T. research a novel coronavirus from patients with pneumonia in China, 2019. New Engl J Med. 2020;382:727–33. https://doi.org/10.1056/NEJMoa2001017.

    Article  CAS  PubMed  Google Scholar 

  101. Mehta P, McAuley DF, Brown M, Sanchez E, Tattersall RS, Manson JJ, U. K. Hlh Across Speciality Collaboration. COVID-19: consider cytokine storm syndromes and immunosuppression. Lancet. 2020;395:1033–4. https://doi.org/10.1016/S0140-6736(20)30628-0.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Verma G, Dhawan M, Saied AA, Kaur G, Kumar R, Emran TB. Immunomodulatory approaches in managing lung inflammation in COVID-19: a double-edge sword. Immun Inflamm Dis. 2023;11:e1020. https://doi.org/10.1002/iid3.1020.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Malemnganba T, Rattan A, Prajapati VK. Decoding macrophage immunometabolism in human viral infection. Adv Protein Chem Struct Biol. 2024;140:493–523. https://doi.org/10.1016/bs.apcsb.2023.12.003.

    Article  PubMed  Google Scholar 

  104. Olagnier D, Farahani E, Thyrsted J, Blay-Cadanet J, Herengt A, Idorn M, Hait A, Hernaez B, Knudsen A, Iversen MB, Schilling M, Jorgensen SE, Thomsen M, Reinert LS, Lappe M, Hoang HD, Gilchrist VH, Hansen AL, Ottosen R, Nielsen CG, Moller C, van der Horst D, Peri S, Balachandran S, Huang J, Jakobsen M, Svenningsen EB, Poulsen TB, Bartsch L, Thielke AL, Luo Y, Alain T, Rehwinkel J, Alcami A, Hiscott J, Mogensen TH, Paludan SR, Holm CK. SARS-CoV2-mediated suppression of NRF2-signaling reveals potent antiviral and anti-inflammatory activity of 4-octyl-itaconate and dimethyl fumarate. Nat Commun. 2020;11:4938. https://doi.org/10.1038/s41467-020-18764-3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Zhang S, Wang J, Wang L, Aliyari S, Cheng G. SARS-CoV-2 virus NSP14 Impairs NRF2/HMOX1 activation by targeting Sirtuin 1. Cell Mol Immunol. 2022;19:872–82. https://doi.org/10.1038/s41423-022-00887-w.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Nguyen V, Zhang Y, Gao C, Cao X, Tian Y, Carver W, Kiaris H, Cui T, Tan W. The spike protein of SARS-CoV-2 impairs lipid metabolism and increases susceptibility to lipotoxicity: implication for a role of Nrf2. Cells. 2022. https://doi.org/10.3390/cells11121916.

    Article  PubMed  PubMed Central  Google Scholar 

  107. Zhao S, Ghosh A, Lo CS, Chenier I, Scholey JW, Filep JG, Ingelfinger JR, Zhang SL, Chan JSD. Nrf2 deficiency upregulates intrarenal angiotensin-converting enzyme-2 and angiotensin 1–7 receptor expression and attenuates hypertension and nephropathy in diabetic mice. Endocrinology. 2018;159:836–52. https://doi.org/10.1210/en.2017-00752.

    Article  CAS  PubMed  Google Scholar 

  108. Ryan TAJ, Hooftman A, Rehill AM, Johansen MD, Brien ECO, Toller-Kawahisa JE, Wilk MM, Day EA, Weiss HJ, Sarvari P, Vozza EG, Schramm F, Peace CG, Zotta A, Miemczyk S, Nalkurthi C, Hansbro NG, McManus G, O’Doherty L, Gargan S, Long A, Dunne J, Cheallaigh CN, Conlon N, Carty M, Fallon PG, Mills KHG, Creagh EM, Donnell JSO, Hertzog PJ, Hansbro PM, McLoughlin RM, Wygrecka M, Preston RJS, Zaslona Z, O’Neill LAJ. Dimethyl fumarate and 4-octyl itaconate are anticoagulants that suppress tissue factor in macrophages via inhibition of type I interferon. Nat Commun. 2023;14:3513. https://doi.org/10.1038/s41467-023-39174-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Lederer DJ, Martinez FJ. Idiopathic pulmonary fibrosis. New Engl J Med. 2018;378:1811–23. https://doi.org/10.1056/NEJMra1705751.

    Article  CAS  PubMed  Google Scholar 

  110. Thannickal VJ, Toews GB, White ES, Lynch JP 3rd, Martinez FJ. Mechanisms of pulmonary fibrosis. Annu Rev Med. 2004;55:395–417. https://doi.org/10.1146/annurev.med.55.091902.103810.

    Article  CAS  PubMed  Google Scholar 

  111. Ogger PP, Albers GJ, Hewitt RJ, O’Sullivan BJ, Powell JE, Calamita E, Ghai P, Walker SA, McErlean P, Saunders P, Kingston S, Molyneaux PL, Halket JM, Gray R, Chambers DC, Maher TM, Lloyd CM, Byrne AJ. Itaconate controls the severity of pulmonary fibrosis. Sci Immunol. 2020;5:eabc1884. https://doi.org/10.1126/sciimmunol.abc1884.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Han YY, Gu X, Yang CY, Ji HM, Lan YJ, Bi YQ, Si R, Qu J, Cheng MH, Gao J. Protective effect of dimethyl itaconate against fibroblast-myofibroblast differentiation during pulmonary fibrosis by inhibiting TXNIP. J Cell Physiol. 2021;236:7734–44. https://doi.org/10.1002/jcp.30456.

    Article  CAS  PubMed  Google Scholar 

  113. Hao W, Li M, Cai Q, Wu S, Li X, He Q, Hu Y. Roles of NRF2 in fibrotic diseases: from mechanisms to therapeutic approaches. Front Physiol. 2022;13:889792. https://doi.org/10.3389/fphys.2022.889792.

    Article  PubMed  PubMed Central  Google Scholar 

  114. Sockrider M, Fussner L. What Is Asthma? Am J Resp Crit Care. 2020;202:P25–6. https://doi.org/10.1164/rccm.2029P25.

    Article  Google Scholar 

  115. Wang L, Chen X, Li X, Liu D, Wang X, Chang X, Guo Y. Developing a novel strategy for COPD therapy by targeting Nrf2 and metabolism reprogramming simultaneously. Free Radic Bio Med. 2021;169:436–45. https://doi.org/10.1016/j.freeradbiomed.2021.03.039.

    Article  CAS  Google Scholar 

  116. Bewley MA, Preston JA, Mohasin M, Marriott HM, Budd RC, Swales J, Collini P, Greaves DR, Craig RW, Brightling CE, Donnelly LE, Barnes PJ, Singh D, Shapiro SD, Whyte MKB, Dockrell DH. Impaired mitochondrial microbicidal responses in chronic obstructive pulmonary disease macrophages. Am J Resp Crit Care. 2017;196:845–55. https://doi.org/10.1164/rccm.201608-1714OC.

    Article  CAS  Google Scholar 

  117. Bade BC, Dela CS. Cruz lung cancer 2020: epidemiology, etiology, and prevention. Clin Chest Med. 2020;41:1–24. https://doi.org/10.1016/j.ccm.2019.10.001.

    Article  PubMed  Google Scholar 

  118. Yang W, Wang Y, Huang Y, Yu J, Wang T, Li C, Yang L, Zhang P, Shi L, Yin Y, Tao K, Li R. 4-Octyl itaconate inhibits aerobic glycolysis by targeting GAPDH to promote cuproptosis in colorectal cancer. Biomed Pharmacother. 2023;159:114301. https://doi.org/10.1016/j.biopha.2023.114301.

    Article  CAS  PubMed  Google Scholar 

  119. Gautam AK, Kumar P, Raj R, Kumar D, Bhattacharya B, Rajinikanth PS, Chidambaram K, Mahata T, Maity B, Saha S. Preclinical evaluation of dimethyl itaconate against hepatocellular carcinoma via activation of the e/iNOS-mediated NF-kappaB-dependent apoptotic pathway. Front Pharmacol. 2021;12:823285. https://doi.org/10.3389/fphar.2021.823285.

    Article  CAS  PubMed  Google Scholar 

  120. Hayashi K, Nakazato Y, Ouchi M, Fujita T, Endou H, Chida M. Antitumor effect of dimethyl itaconate on thymic carcinoma by targeting LDHA-mTOR axis. Life Sci. 2021;282:119847. https://doi.org/10.1016/j.lfs.2021.119847.

    Article  CAS  PubMed  Google Scholar 

  121. Wang J, Lu Q, Cai J, Wang Y, Lai X, Qiu Y, Huang Y, Ke Q, Zhang Y, Guan Y, Wu H, Wang Y, Liu X, Shi Y, Zhang K, Wang M, Peng Xiang A. Nestin regulates cellular redox homeostasis in lung cancer through the Keap1-Nrf2 feedback loop. Nat Commun. 2019;10:5043. https://doi.org/10.1038/s41467-019-12925-9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Scheurlen KM, Snook DL, Walter MN, Cook CN, Fiechter CR, Pan J, Beal RJ, Galandiuk S. Itaconate and leptin affecting PPARgamma in M2 macrophages: a potential link to early-onset colorectal cancer. Surgery. 2022;171:650–6. https://doi.org/10.1016/j.surg.2021.10.054.

    Article  PubMed  Google Scholar 

  123. Zhao H, Teng D, Yang L, Xu X, Chen J, Jiang T, Feng AY, Zhang Y, Frederick DT, Gu L, Cai L, Asara JM, Pasca di Magliano M, Boland GM, Flaherty KT, Swanson KD, Liu D, Rabinowitz JD, Zheng B. Myeloid-derived itaconate suppresses cytotoxic CD8(+) T cells and promotes tumour growth. Nat Metab. 2022;4:1660–73. https://doi.org/10.1038/s42255-022-00676-9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Zhao Y, Liu Z, Liu G, Zhang Y, Liu S, Gan D, Chang W, Peng X, Sung ES, Gilbert K, Zhu Y, Wang X, Zeng Z, Baldwin H, Ren G, Weaver J, Huron A, Mayberry T, Wang Q, Wang Y, Diaz-Rubio ME, Su X, Stack MS, Zhang S, Lu X, Sheldon RD, Li J, Zhang C, Wan J, Lu X. Neutrophils resist ferroptosis and promote breast cancer metastasis through aconitate decarboxylase 1. Cell Metab. 2023;35:1688–703. https://doi.org/10.1016/j.cmet.2023.09.004.

    Article  CAS  PubMed  Google Scholar 

  125. Wang X, Su S, Zhu Y, Cheng X, Cheng C, Chen L, Lei A, Zhang L, Xu Y, Ye D, Zhang Y, Li W, Zhang J. Metabolic reprogramming via ACOD1 depletion enhances function of human induced pluripotent stem cell-derived CAR-macrophages in solid tumors. Nat Commun. 2023;14:5778. https://doi.org/10.1038/s41467-023-41470-9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. ElAzzouny M, Tom CT, Evans CR, Olson LL, Tanga MJ, Gallagher KA, Martin BR, Burant CF. Dimethyl itaconate is not metabolized into itaconate intracellularly. J Biol Chem. 2017;292:4766–9. https://doi.org/10.1074/jbc.C117.775270.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Chen F, Elgaher WAM, Winterhoff M, Bussow K, Waqas FH, Graner E, Pires-Afonso Y, Casares Perez L, de la Vega L, Sahini N, Czichon L, Zobl W, Zillinger T, Shehata M, Pleschka S, Bahre H, Falk C, Michelucci A, Schuchardt S, Blankenfeldt W, Hirsch AKH, Pessler F. Citraconate inhibits ACOD1 (IRG1) catalysis, reduces interferon responses and oxidative stress, and modulates inflammation and cell metabolism. Nat Metab. 2022;4:534–46. https://doi.org/10.1038/s42255-022-00577-x.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Li Z, Zheng W, Kong W, Zeng T. Itaconate: a potent macrophage immunomodulator. Inflammation. 2023;46:1177–91. https://doi.org/10.1007/s10753-023-01819-0.

    Article  CAS  PubMed  Google Scholar 

  129. Liu D, Xiao W, Li H, Zhang Y, Yuan S, Li C, Dong S, Wang C. Discovery of itaconate-mediated lysine acylation. J Am Chem Soc. 2023;145:12673–81. https://doi.org/10.1021/jacs.3c02332.

    Article  CAS  PubMed  Google Scholar 

  130. Hu Z, Xu D, Meng H, Liu W, Zheng Q, Wang J. 4-octyl itaconate protects against oxidative stress-induced liver injury by activating the Nrf2/Sirt3 pathway through AKT and ERK1/2 phosphorylation. Biochem Pharmacol. 2024;220:115992. https://doi.org/10.1016/j.bcp.2023.115992.

    Article  CAS  PubMed  Google Scholar 

  131. Wang X, Kong W, Yang R, Yang C. 4-octyl itaconate ameliorates ventilator-induced lung injury. Arch Biochem Biophys. 2024;752:109853. https://doi.org/10.1016/j.abb.2023.109853.

    Article  CAS  PubMed  Google Scholar 

  132. Ryan DG, Murphy MP, Frezza C, Prag HA, Chouchani ET, O’Neill LA, Mills EL. Coupling Krebs cycle metabolites to signalling in immunity and cancer. Nat Metab. 2019;1:16–33. https://doi.org/10.1038/s42255-018-0014-7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Chen X, Li C, Cao X, Jia X, Chen X, Wang Z, Xu W, Dai F, Zhang S. Mitochondria-targeted supramolecular coordination container encapsulated with exogenous itaconate for synergistic therapy of joint inflammation. Theranostics. 2022;12:3251–72. https://doi.org/10.7150/thno.70623.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Riquelme SA, Liimatta K, Lung TWF, Fields B, Ahn D, Chen D, Lozano C, Saenz Y, Uhlemann AC, Kahl BC, Britto CJ, DiMango E, Prince A. Pseudomonas aeruginosa utilizes host-derived itaconate to redirect its metabolism to promote biofilm formation. Cell Metab. 2020;31:1091–106. https://doi.org/10.1016/j.cmet.2020.04.017.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Li Y, Chen X, Zhang H, Xiao J, Yang C, Chen W, Wei Z, Chen X, Liu J. 4-Octyl itaconate alleviates lipopolysaccharide-induced acute lung injury in mice by inhibiting oxidative stress and inflammation. Drug Des Dev Ther. 2020;14:5547–58. https://doi.org/10.2147/DDDT.S280922.

    Article  CAS  Google Scholar 

  136. Yang W, Wang Y, Huang Y, Wang T, Li C, Zhang P, Liu W, Yin Y, Li R, Tao K. Immune Response Gene-1 [IRG1]/itaconate protect against multi-organ injury via inhibiting gasdermin D-mediated pyroptosis and inflammatory response. Inflammopharmacology. 2024;32:419–32. https://doi.org/10.1007/s10787-023-01278-x.

    Article  CAS  PubMed  Google Scholar 

  137. Zeng YR, Song JB, Wang D, Huang ZX, Zhang C, Sun YP, Shu G, Xiong Y, Guan KL, Ye D, Wang P. The immunometabolite itaconate stimulates OXGR1 to promote mucociliary clearance during the pulmonary innate immune response. J Clin Invest. 2023. https://doi.org/10.1172/JCI160463.

    Article  PubMed  PubMed Central  Google Scholar 

  138. Tomlinson KL, Riquelme SA, Baskota SU, Drikic M, Monk IR, Stinear TP, Lewis IA, Prince AS. Staphylococcus aureus stimulates neutrophil itaconate production that suppresses the oxidative burst. Cell Rep. 2023;42:112064. https://doi.org/10.1016/j.celrep.2023.112064.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Demars A, Vitali A, Comein A, Carlier E, Azouz A, Goriely S, Smout J, Flamand V, Van Gysel M, Wouters J, Abendroth J, Edwards TE, Machelart A, Hoffmann E, Brodin P, De Bolle X, Muraille E. Aconitate decarboxylase 1 participates in the control of pulmonary Brucella infection in mice. PLoS Pathog. 2021;17:e1009887. https://doi.org/10.1371/journal.ppat.1009887.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Sun KA, Li Y, Meliton AY, Woods PS, Kimmig LM, Cetin-Atalay R, Hamanaka RB, Mutlu GM. Endogenous itaconate is not required for particulate matter-induced NRF2 expression or inflammatory response. Elife. 2020. https://doi.org/10.7554/eLife.54877.

    Article  PubMed  PubMed Central  Google Scholar 

  141. Wu YX, Zhang YR, Jiang FJ, He S, Zhang YL, Chen D, Tong Y, Nie YJ, Pang QF. 4-OI ameliorates bleomycin-induced pulmonary fibrosis by activating Nrf2 and suppressing macrophage-mediated epithelial-mesenchymal transition. Inflamm Res. 2023;72:1133–45. https://doi.org/10.1007/s00011-023-01733-z.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We acknowledge TopEdit LLC for the linguistic editing and proofreading during the preparation of this manuscript.

Funding

This work was supported by grants from the National Natural Science Foundation of China (no. 8210082163 and 82300103), the Fundamental Research Funds for the Central Universities (no. 2042021kf0081 and 2042023kf0011) and the Science Fund for Creative Research Groups of the Natural Science Foundation of Hubei Province (No.2020CFA027).

Author information

Authors and Affiliations

Authors

Contributions

R.H., Y.Z., and K.Y. wrote the main manuscript text, and B.L. provided significant input. S.C., N.L., and Q.G. reviewed and edited the manuscript and approved the final version. All authors have read and approved the manuscript.

Corresponding authors

Correspondence to Congkuan Song, Ning Li or Qing Geng.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

He, R., Zuo, Y., Yi, K. et al. The role and therapeutic potential of itaconate in lung disease. Cell Mol Biol Lett 29, 129 (2024). https://doi.org/10.1186/s11658-024-00642-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s11658-024-00642-1

Keywords