Skip to main content
Journal of Asthma and Allergy logoLink to Journal of Asthma and Allergy
. 2025 Jul 25;18:1129–1146. doi: 10.2147/JAA.S529371

Reactive Oxygen Species in Asthma: Regulators of Macrophage Polarization and Therapeutic Implications: A Narrative Review

Ying Liu 1, Mingyao Zhang 2, Tongtong Wang 3, Jun Zhang 1,
PMCID: PMC12306563  PMID: 40735422

Abstract

As a vital component of the immune system, macrophages play a critical role in the progression of asthma. The two classic polarization states of macrophages, M1 and M2, exhibit distinct functions. M1-polarized macrophages eliminate pathogens through the secretion of pro-inflammatory cytokines, while M2-polarized macrophages secrete anti-inflammatory factors to facilitate tissue repair. However, in asthma, the activation of M1 macrophages is often associated with excessive inflammatory responses, whereas M2 macrophages contribute to airway remodeling and chronic inflammation. These processes collectively exacerbate airway inflammation and remodeling, thereby aggravating asthma symptoms. Reactive oxygen species (ROS), as crucial signaling molecules, have been shown to regulate macrophage polarization and promote both M1 and M2 polarization states. This review summarizes the primary endogenous and exogenous sources of ROS in asthma and elaborates on the mechanisms by which ROS influence M1/M2 polarization of macrophages. Endogenous ROS arise chiefly from NOX2, xanthine oxidase, peroxisomes and mitochondria, whereas ozone and fine particulate matter are major exogenous sources. ROS activate MAPK, NF-κB and NLRP3 cascades, boosting IL-1β, IL-6 and IL-27 release by M1 cells, while low NOX2 flux or mitochondrial H2O2 supports STAT6-dependent ARG1 expression and drives an M2 program. Additionally, we discuss the impact of different macrophage polarization states on asthma pathophysiology and the potential applications of macrophage-modulating agents in asthma treatment, particularly those targeting ROS-mediated polarization pathways. ARG1 rich M2 cells convert L-arginine into proline, fostering collagen deposition; Ym1/2, Fizz1 and CD206 correlate with airway remodeling and declining lung function. Emerging antioxidant and macrophage-polarization strategies that selectively modulate ROS show promise in rebalancing M1/M2 responses and attenuating airway hyper-responsiveness. This review provides new insights into the interplay between ROS and macrophage polarization and highlights the potential for developing therapies aimed at modulating macrophage polarization via ROS regulation.

Keywords: asthma, macrophage polarization, reactive oxygen species

Introduction

Asthma is a chronic inflammatory airway disease characterized by recurrent wheezing, shortness of breath, chest tightness, and coughing.1 In recent years, the prevalence of asthma has risen significantly, affecting nearly 250 million people worldwide and causing over 1000 deaths daily.2 Despite significant advancements in asthma treatment over the past 15 years, which have allowed most patients to achieve effective disease control, the underlying etiology of asthma remains unclear.3 Studies suggest that both environmental and genetic factors play critical roles in asthma pathogenesis. Genetic factors influence individual sensitivity to allergens, while environmental factors, such as allergen exposure, trigger allergic inflammation, a hallmark pathological feature of asthma.4,5 In Indian rhesus macaques, macrophages make up about 70% of lung immune cells and play a pivotal role in airway inflammation induced by allergens.6

Pulmonary macrophages primarily consist of alveolar macrophages (AMs) and interstitial macrophages. AMs represent the majority (75–80%) and exhibit a low turnover rate, whereas interstitial macrophages display a higher turnover rate and share similarities with circulating monocytes, suggesting that monocytes might be a key source of interstitial macrophages.6,7 The polarization state of macrophages is a crucial determinant of immune response. As shown in Figure 1, macrophages can polarize into two major phenotypes: classically activated (M1) and alternatively activated (M2). M1 macrophages typically exhibit pro-inflammatory functions, producing inflammatory cytokines and reactive oxygen species (ROS) to eliminate pathogens. In contrast, M2 macrophages are associated with anti-inflammatory responses, tissue repair, and immune suppression. A significant increase in M2 macrophages has been observed in asthma patients, where they contribute to airway remodeling and chronic inflammation through the secretion of cytokines such as IL-13, IL-6, and CCL2.8 Therefore, dysregulated macrophage polarization is implicated in airway inflammation and exacerbation of asthma symptoms.

Figure 1.

Figure 1

M1/M2 Polarization States of Macrophages and Their Functions. M1-polarized macrophages secrete pro-inflammatory cytokines and produce high levels of reactive oxygen species (ROS) to eliminate pathogens. In asthma, excessively activated M1 macrophages contribute to airway inflammation. M2-polarized macrophages, on the other hand, secrete anti-inflammatory cytokines, suppress immune activation, and promote tissue repair. Overactivation of M2 macrophages during early asthma leads to airway remodeling, which ultimately results in airway obstruction. Created in Adobe Illustrator.

ROS, highly reactive oxygen-containing molecules, have been extensively studied for their role in promoting asthma progression.9 In the asthmatic airway, inflammatory cells are recruited and produce various ROS, driving tissue damage and chronic airway inflammation.10 Moreover, ROS actively participate in immune regulatory responses, including the modulation of macrophage polarization.11 Understanding the sources of ROS in asthma and how ROS influence macrophage polarization not only sheds light on the immunopathology of asthma but also identifies potential therapeutic targets for disease management.

This review focuses on the role of ROS in asthma, particularly in regulating macrophage polarization. We will summarize how ROS influence immune cell functions, especially macrophage polarization, to modulate airway inflammation and immune responses. Furthermore, we will explore the potential therapeutic applications of ROS scavengers in asthma immunotherapy, offering new insights for future treatment strategies.

Sources of ROS in Asthma

ROS can be classified into endogenous and exogenous sources (Figure 2). Endogenous ROS are primarily generated by oxidases, including NADPH oxidase (NOX), xanthine oxidase, and peroxisomes, as well as by mitochondria. Exogenous ROS, on the other hand, are primarily derived from environmental pollutants.

Figure 2.

Figure 2

Sources of reactive oxygen species. ROS originate from endogenous (A) and exogenous (B) sources. (A) Endogenous ROS are produced mainly by oxidases, notably NADPH oxidase (NOX2), xanthine oxidase, and the peroxisomal β oxidation system, together with the mitochondrial electron transport chain. Within NOX2 the transmembrane subunits gp91phox and p22phox form its catalytic core. The cytosolic subunits p40phox, p47phox, and p67phox, assisted by the small GTPases Rac1 and Rac2, move to the membrane after phosphorylation and guanine nucleotide exchange, stabilising the complex; electrons then flow from cytosolic NADPH to molecular oxygen and generate superoxide. Xanthine oxidase converts hypoxanthine and xanthine to uric acid while releasing superoxide and hydrogen peroxide and can further enhance mitochondrial ROS through the AKT PI3K mTOR pathway. Peroxisomes liberate H2O2 during fatty-acid β oxidation. In mitochondria electrons that leak from complex I and complex III, and under stress also from complex IV, reduce O2 to superoxide, which is subsequently dismutated to H2O2. (B) Exogenous ROS arise mainly from air pollutants such as ozone and particulate matter. Tropospheric ozone is generated photochemically from nitrogen oxides and volatile organic compounds in vehicle exhaust. Although antioxidants in the epithelial lining fluid, including glutathione, ascorbic acid, vitamin E, and uric acid, react with ozone, high ozone concentrations can overwhelm this defence and promote ROS buildup. Particulate matter contains polycyclic aromatic hydrocarbons that undergo redox cycling to form superoxide, and fine particles rich in transition metal ions catalyse hydroxyl-radical formation through the Fenton reaction, thereby amplifying oxidative stress. Created in Adobe Illustrator.

Endogenous Sources

NADPH Oxidase

NADPH oxidase is a family of enzymes that utilize nicotinamide adenine dinucleotide phosphate (NADPH) as an electron donor to catalyze the reduction of O2 into superoxide anion (O2•−).12 The NADPH oxidase family comprises several members, including NOX1 through NOX5.13 Each member plays distinct roles in different cell types and physiological processes. Among these, NOX2 is primarily expressed in phagocytic cells such as eosinophils, neutrophils, macrophages, and dendritic cells.14 NOX2 is a multi-subunit protein complex consisting of two transmembrane proteins, gp91phox and p22phox, and three cytosolic components, p40phox, p47phox, and p67phox.13 The catalytic core of NOX2 is formed by gp91phox and p22phox, while p40phox, p47phox, and p67phox assemble into a cytosolic trimer that binds to the catalytic core, forming the functional NOX2 complex.15 This assembly process is regulated by the activation of G protein-coupled receptors, Fc receptors, and integrin receptors, which phosphorylate p47phox and p67phox via protein kinase C (PKC). The phosphorylation of p47phox and p67phox induces conformational changes that promote their translocation to the membrane.16,17 Toll-like receptors (TLRs) also contribute to NOX2 activation, although not directly. Instead, TLRs partially phosphorylate p47phox, priming cells for heightened sensitivity to subsequent stimuli.18 Additionally, the small GTPases Rac1 and Rac2 play crucial roles in NOX2 activation. Under the influence of guanine nucleotide exchange factors (GEFs), Rac1 and Rac2 bind to GTP and translocate from the cytosol to the membrane. There, they interact with the catalytic core of NOX2, stabilizing the complex and facilitating efficient electron transfer.16,19 Within the catalytic core of the NOX2 complex, electrons are transferred from intracellular NADPH through the transmembrane region (formed by gp91phox and p22phox). These electrons are ultimately delivered to extracellular oxygen molecules, producing O2•−.20

Xanthine Oxidase

Xanthine oxidase (XO) is a member of the molybdenum-flavin enzyme family and contains molybdenum and flavin as cofactors.21 XO typically exists as a homodimer, with each subunit comprising an Fe-S redox-active center, a flavin cofactor, and a substrate-binding domain.22 Predominantly localized in the cytoplasm, XO catalyzes the oxidation of hypoxanthine and xanthine into uric acid, generating O2•− and hydrogen peroxide (H2O2) as byproducts.23 This property makes XO a potential source of ROS.24 Studies have shown that the use of allopurinol, an XO inhibitor, in asthmatic mice suppresses airway inflammation and Th1, Th2, and Th17 immune responses, thereby mitigating airway damage.9 ROS produced by XO can activate the p38 mitogen activated protein kinase (MAPK)- nuclear factor of activated T cells 5 (NFAT5) signaling pathway and promote the expression of the pro-inflammatory cytokine IL-6, a hallmark of M1 macrophages. This suggests that XO may contribute to the polarization of macrophages toward the M1 phenotype.25 Research by Annette Ives et al has demonstrated that XO is a major source of ROS in macrophages. XO activates the AKT-PI3K-mTOR pathway, enhancing mitochondrial ROS (mtROS) production and simultaneously activating the NLRP3 inflammasome in mice. This dual mechanism drives macrophages toward a pro-inflammatory M1 phenotype.26

Peroxisomes

Peroxisomes are cytoplasmic organelles widely present in nearly all eukaryotic cells. They play a crucial role in the oxidative metabolism of amino acids and the β-oxidation of fatty acids, during which ROS are primarily generated in the form of H2O2 and O2•−.27,28 L-amino acid oxidase, D-amino acid oxidase, and polyamine oxidase catalyze the oxidation of amino acids, transferring electrons to O2 and releasing H2O2 as a byproduct.29,30 Fatty acids are broken down into acetyl-CoA via β-oxidation in peroxisomes and subsequently transferred to mitochondria to participate in the tricarboxylic acid (TCA) cycle. During this process, flavin adenine dinucleotide (FAD) in acyl-CoA oxidase accepts electrons from acyl-CoA and transfers them to O2, resulting in the production of H2O2.31 The H2O2 generated in peroxisomes can be degraded by catalase and glutathione peroxidase, enzymes contained within the organelle. However, an excessive influx of fatty acids into peroxisomes can lead to the accumulation of H2O2, causing elevated ROS levels and subsequent oxidative stress.32

Mitochondria

Mitochondria, as the primary energy suppliers of cells, generate mtROS during respiration, with O2•− being the predominant form.33 In mitochondria, NADH and reduced flavin adenine dinucleotide (FADH2) transfer high-energy electrons to O2 via the electron transport chain (ETC). This process produces ROS while simultaneously driving proton pumping from the mitochondrial matrix into the intermembrane space, creating a proton gradient that powers ATP synthesis through ATP synthase.34 mtROS are primarily generated at Complexes I through IV of the ETC, with Complexes I and III being the major contributors.35 Complex I serves as the entry point of electrons from NADH into the ETC.36 The flavin mononucleotide (FMN) cofactor accepts electrons from NADH and transfers them to coenzyme Q (CoQ). During this transfer, FMN may leak electrons to O2, producing O2•−.37 Complex III transfers electrons from CoQ to cytochrome c and features two CoQ binding sites: the Qi and Qo sites. At the Qo site, CoQ is oxidized, releasing electrons that are subsequently transferred to cytochrome c and the Qi site for CoQ reduction.38 When the Qi site is inhibited, electrons accumulate at Complex III and are transferred to O2 via the semiquinone intermediate at the Qo site, producing O2•−.39 These O2•− are released to both sides of the mitochondrial inner membrane via Complex III40 and are subsequently converted into H2O2 by mitochondrial superoxide dismutase (SOD).41 O2•− in the intermembrane space can also reach the cytoplasm through voltage-dependent anion channels.42 In addition to Complexes I and III, Complexes II and IV have also been implicated in mtROS production, though their mechanisms are less well understood.43 The FAD cofactor in Complex II is thought to be a major site of mtROS generation.43 Complex IV, as the terminal point of the ETC, rarely leaks electrons under normal conditions. However, under pathological conditions such as mitochondrial damage, partially reduced oxygen species may escape from Complex IV, forming ROS.44

Research suggests that the activation of innate immune signaling recruits mitochondria to macrophage phagosomes, enhancing mtROS production and increasing macrophage bactericidal activity in mice.45 Lipopolysaccharide (LPS) stimulation induces metabolic reprogramming in macrophages, shifting from oxidative phosphorylation to glycolysis, which promotes mtROS production at Complex I and enhances the expression of M1 polarization markers such as IL-1β.46 This may be attributed to mtROS-mediated oxidation of succinate dehydrogenase, stabilizing hypoxia-inducible factor 1-alpha (HIF-1α) and subsequently driving sustained IL-1β expression.47 Furthermore, mtROS can activate the p38 MAPK signaling pathway, stimulating the expression of pro-inflammatory cytokines.48

Exogenous Sources: Air Pollutants as ROS Generators

Air pollutants significantly impact the onset of asthma and exacerbate its symptoms.49 Traffic-related pollutants, such as ozone (O₃) and particulate matter (PM), are potent oxidants that contribute to ROS production.50

Ozone

O₃, a common air pollutant, is primarily formed through photochemical reactions involving nitrogen oxides and volatile organic compounds from vehicle exhaust.51 Studies indicate that the airway epithelial lining fluid (ELF) serves as the first line of defense against O₃ exposure. Antioxidants present in ELF, such as glutathione (GSH), ascorbic acid, vitamin E, and uric acid, react with and neutralize most of the O₃.52 However, high concentrations of O₃ can overwhelm the antioxidant capacity of the ELF, leading to excessive ROS generation and the release of inflammatory cytokines like IL-33.53 In allergic asthma, epithelial-barrier integrity is frequently impaired, making airway epithelial cells particularly vulnerable to ROS released by infiltrating inflammatory cells. These ROS activate NF-κB and markedly amplify ozone-induced IL-8 production.54 Neutrophils from asthmatic patients, in turn, display an exaggerated responsiveness to IL-8.55 The chemokine also engages CXCR1/CXCR2 receptors on airway smooth-muscle cells, provoking bronchoconstriction56 and further exacerbating asthma.55

Particulate Matter

PM, a major component of air pollution, primarily originates from vehicle emissions, industrial discharges, and construction activities. PM with a diameter of less than 10 micrometers (PM₁0) is strongly associated with increased incidence and mortality of respiratory diseases.57 Interestingly, the oxidative potential of PM, rather than PM itself, has shown a stronger correlation with adverse health effects.58 PM contains high levels of polycyclic aromatic hydrocarbons (PAHs), which undergo metabolic activation by cytochrome P450 enzymes to form epoxides. These epoxides are converted into dihydrodiol intermediates by epoxide hydrolases, which are subsequently oxidized into quinones by dihydrodiol dehydrogenases.59,60 Quinones are then reduced to semiquinone radicals by NADPH cytochrome P450 reductase. These semiquinone radicals are unstable and rapidly transfer electrons to oxygen, generating O2•−.61 Different PM sizes impose varying oxidative stress burdens. Smaller PM particles tend to contain higher levels of transition metal ions, such as nickel (Ni), chromium (Cr), and iron (Fe), which catalyze the production of hydroxyl radicals through Fenton reactions.62 Once PM enters the airways, it is phagocytosed by macrophages, which subsequently promote inflammation and exacerbate disease progression.59

Regulation of Macrophage Polarization by ROS

The process by which macrophages exhibit distinct functional phenotypes in response to signals within their microenvironment is known as macrophage polarization.63 Although recent studies suggest that macrophage polarization involves diverse and partially overlapping transcriptional programs, the classical M1/M2 paradigm remains widely used to reflect macrophage functional states.64 Macrophages with a pro-inflammatory phenotype are categorized as M1 macrophages, while those with anti-inflammatory or immunosuppressive phenotypes are classified as M2 macrophages.65 At the molecular level, M1 macrophages are characterized by the expression of pro-inflammatory cytokines, such as TNF-α, IL-1β, IL-6, and IL-12, whereas M2 macrophages are associated with the expression of cytokines like TGF-β, CCL18, and IL-1Ra.66 Different macrophage polarization states play distinct roles in asthma. As shown in Figure 3, recent research has shown that ROS act as regulatory factors capable of influencing and modulating the M1/M2 polarization of macrophages.67,68

Figure 3.

Figure 3

The Role of ROS in Macrophage Polarization. Under IFN-γ and LPS stimulation, corresponding receptors are activated, initiating the STAT1 and MAPK pathways to drive M1 macrophage polarization, a process that depends on ROS. ROS also facilitate M1 polarization by promoting ASK phosphorylation and activating NF-κB. Additionally, mtROS are essential for NLRP3 inflammasome-dependent M1 polarization. When macrophages engulf apoptotic materials, SYNCRIP expression is downregulated, leading to reduced stability of NOX2 mRNA, decreased NOX2 protein levels, and lower ROS production. The absence of the p47phox subunit in NOX2 has been shown to increase sensitivity to external stimuli, thereby enhancing Stat6 phosphorylation and promoting M2 polarization. Furthermore, mitochondrial ROS and H2O2 contribute to M2 macrophage polarization by activating non-canonical NF-κB and enhancing Stat6 phosphorylation, respectively. Created in Adobe Illustrator.

ROS in Driving M1 Macrophage Polarization

ROS play a critical role in driving M1 polarization of macrophages. The surface receptor IFN-γR on macrophages recognizes IFN-γ secreted by immune cells, leading to the activation of Janus kinase (JAK).69 Phosphorylated signal transducer and activator of transcription (STAT)1, activated by JAK, translocates into the nucleus and regulates the transcription of M1-related genes, such as TNF-α, inducible nitric oxide synthase (iNOS), and IL-12.70 The LPS/TLR4 pathway is also crucial for M1 polarization.71 Binding of LPS to TLR4 recruits myeloid differentiation primary response 88 (MyD88) and TIR-domain-containing adapter-inducing interferon-β (TRIF).72 MyD88 activates the MAPK signaling pathway, promoting the expression of inflammatory cytokines and enzymes like iNOS.73 MyD88 also triggers the nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) pathway, a key event in M1 polarization, enabling NF-κB to translocate into the nucleus and drive the expression of pro-inflammatory cytokines such as IL-1β, IL-6, and TNF-α.74 In addition, MyD88 has been reported to activate the STAT1 pathway, further promoting M1 polarization.75,76 MyD88 also activates the transcription factor IRF5, which directly induces the genes encoding IL-12 subunits p40, IL-12p35, and IL-23p19, thereby enhancing M1 polarization.77,78 The TRIF pathway, on the other hand, facilitates the production of type I interferons, amplifying M1-associated transcriptional programs.79 Upon stimulation by IFN-γ and LPS, rapid activation of Phox and ROS release have been observed in microglia and astrocytes. The use of Phox inhibitors and catalase can suppress the activation of MAPK, NF-κB, and JAK/STAT pathways under IFN-γ and LPS stimulation, reducing the expression of M1 markers such as IL-1, IL-6, TNF-α, and iNOS.80 This suggests that ROS may function as signaling molecules in the activation of the LPS/TLR4 pathway. Activation of the TLR4 pathway has also been shown to stimulate NOX and SOD, leading to the production of H2O2.81 H2O2 can dissociate thioredoxin (Trx) from the apoptosis signal-regulating kinase 1 (ASK1)-Trx complex, inducing homodimerization and phosphorylation of ASK1. Phosphorylated ASK1 subsequently activates downstream p38 MAPK.82 Moreover, ROS enhance the phosphorylation of IκB, marking it for ubiquitination and degradation, which releases and activates NF-κB dimers.81 Recent studies have reported that checkpoint kinase 2 (CHK2) responds to ROS produced under LPS and IFN-γ stimulation by phosphorylating pyruvate kinase M2 (PKM2). Phosphorylated PKM2 promotes glycolysis, induces p21 accumulation, and triggers G1 cell cycle arrest, thereby facilitating M1 polarization.68

The NOD-like receptor family pyrin domain containing 3 (NLRP3) inflammasome senses endogenous damage-associated molecular patterns (DAMPs) and exogenous pathogen-associated molecular patterns (PAMPs), initiating immune responses. It drives the caspase-1-dependent maturation and secretion of IL-1β and IL-18, promoting M1 polarization.83 Numerous studies have indicated that ROS are involved in the activation of the NLRP3 inflammasome.84,85 Targeting mitochondria with MitoQ to block mtROS production has been shown to effectively suppress NLRP3 inflammasome activation and reduce the maturation of IL-1β and IL-18.86 However, cells lacking the p22phox subunit still produce normal levels of IL-1β, suggesting that NLRP3 activation depends on mtROS rather than NADPH oxidase-derived ROS.87

ROS in Promoting M2 Macrophage Polarization

ROS play a dual role in macrophage polarization, not only promoting M1 polarization but also critically regulating M2 polarization. Compared to M1 macrophages, M2 activation is often associated with reduced ROS production and increased arginase-1 (ARG1) activity. M2 macrophages improve phagolysosomal proteolytic activity, facilitating tissue remodeling by reducing NOX2 activity and increasing protease activity.88 The decrease in NOX2 levels in M2 macrophages may result from macrophage interaction with apoptotic cell debris, which downregulates synaptotagmin-binding cytoplasmic RNA-interacting protein (SYNCRIP) expression. This reduces the binding of SYNCRIP to the 3′ UTR of NOX2 mRNA, ultimately lowering NOX2 mRNA abundance and expression, decreasing ROS generation, and promoting M2 polarization.89 Similarly, Padgett et al demonstrated that NOX deficiency limits ROS production, leading to a significant increase in M2 macrophages along with reduced expression of TNF-α and IL-1β.90 The absence of p47phox, a cytosolic subunit of the NOX2 complex, biases macrophages toward M2 polarization. This is due to the heightened sensitivity of p47phox-deficient macrophages to IL-4, resulting in significantly elevated phosphorylation of Stat6 upon IL-4 stimulation.91 This finding suggests that p47phox not only participates in NOX2 complex assembly but also serves as a regulatory factor in the IL-4/Stat6 pathway. Similarly, microglia lacking p47phox exhibit enhanced M2 polarization and increased expression of M2 markers, such as IL-4Rα, chitinase 3-like protein 3 (Ym1), found in inflammatory zone 1 (Fizz1), mannose receptor c-type 1 (Mrc1), cluster of differentiation 163 (CD163), and macrophage receptor with collagenous structure (MARCO), under LPS stimulation. Treating macrophages with the NOX2 inhibitor apocynin produces similar effects.92 Although these studies support the inhibitory role of NOX2-derived ROS in M2 polarization, Zhang et al demonstrated that ROS are essential during the early activation phase of biphasic ERK signaling. Using the NOX2 inhibitor BHA disrupts the differentiation of monocytes into M2 macrophages.93 This suggests that ROS may have stage-specific effects on macrophage polarization, with NOX2-derived ROS being indispensable during the initial activation of M2 macrophages.

Beyond NOX, mtROS also play a critical role in regulating macrophage polarization. mtROS are indispensable for M2 polarization, as studies have shown that inhibiting mtROS in RAW 264.7 cells reduce the expression of the M2 marker ARG1.94 Similar studies suggest that mtROS may activate NF-κB via non-canonical pathways, promoting an anti-inflammatory M2 phenotype.67 The primary pathway for M2 polarization is the Stat6 pathway. Cytokines IL-4 and IL-13 bind to surface receptors, activating JAK, which phosphorylates the intracellular domain of the receptor. The phosphorylated receptor recruits Stat6, which is subsequently phosphorylated by JAK. Activated P-Stat6 forms dimers that translocate to the nucleus, initiating the transcription of M2-related genes.95 Mitochondrial Cu, Zn-SOD converts O2•− into H2O2, and studies have shown that increased Cu, Zn-SOD expression promotes M2 polarization by generating H2O2, which activates STAT6.96

Macrophage Polarization and Its Role in Asthma

In the early stages of asthma, monocyte-derived macrophages are activated and proliferate to facilitate pathogen clearance, initiate pulmonary inflammation, and contribute to the resolution of later-stage inflammation. Studies have shown that during the initial phase of lung injury, circulating Ly6C(hi) monocytes increase significantly, with their numbers positively correlated with the severity of pulmonary inflammation. Conversely, M2-like AMs become predominant in the later stages of the disease and are positively associated with the severity of lung fibrosis.97 Based on the mechanisms of allergic responses, asthma is traditionally categorized into atopic and non-atopic asthma. Atopic asthma is associated with IgE-mediated allergic reactions, whereas non-atopic asthma does not involve IgE-mediated pathways.98 In both atopic and non-atopic asthma, macrophage numbers are elevated. To determine the dominant macrophage phenotypes in these two asthma types, Robbe et al studied a house dust mite (HDM)-induced allergic mouse model and a farm dust extract (FDE)-induced non-allergic mouse model. Their findings revealed that in HDM-induced allergic asthma, macrophages predominantly exhibit M2 polarization, accompanied by a Th2 cell response. In contrast, macrophages in FDE-induced non-allergic asthma primarily display M1 polarization, alongside increased expression of Th1 and Th17 cells.99

M1 Polarization: Inflammation and Pathogenesis in Asthma

M1 macrophages play a vital role in host defense against pathogens by releasing large amounts of pro-inflammatory cytokines and chemokines through phagocytosis.100 However, some studies suggest that bacterial and viral pathogens observed in the airways of atopic asthma patients are insufficient to be effectively controlled by M1-activated macrophages.100,101 Despite this, pro-inflammatory mediators released by M1 macrophages have been shown to exacerbate lung damage and accelerate airway remodeling. For example, elevated levels of nitric oxide (NO) in the exhaled breath of asthma patients are primarily derived from macrophages. High concentrations of NO can induce oxidative DNA damage, trigger inflammation, and enhance mucus secretion in allergen-induced asthma mouse models.102,103 Additionally, M1 macrophages secrete pro-inflammatory cytokines such as IL-1α, IL-1β, IL-6, and TNF-α, which activate CD4+ T cells and enhance IL-5 secretion.104 IL-5 is a major activator of eosinophils, promoting their proliferation, migration, and degranulation, thereby exacerbating airway inflammation and hyperresponsiveness in asthma patients.105 Moreover, IL-6 contributes to pulmonary fibrosis by promoting fibroblast proliferation, activating the MAPK signaling pathway, and synergizing with TGF-β to drive fibrotic progression.106 Animal studies have shown that high levels of LPS can induce type 1 asthma, characterized by non-eosinophilic inflammation, airway hyperresponsiveness, and increased IL-12 expression.107 Administering LPS to the lungs of asthmatic mice increases inflammatory macrophage subtypes (CD11bhighF4/80highCD11cvar) and enhances the expression of inflammatory cytokine IL-27.107 Elevated levels of IL-27 and IFN-γ have been observed in steroid-resistant asthma patients. Their synergistic action inhibits glucocorticoid receptor (GR) nuclear translocation via the MyD88-dependent signaling pathway, contributing to glucocorticoid resistance.108 IL-27 has been reported to activate IFN-inducible gene expression in monocyte-derived cells109 and induce pro-inflammatory gene expression through STAT1-dependent pathways. Additionally, IL-27 amplifies inflammation by enhancing monocyte responses to TLR stimulation and suppressing IL-10 signaling and production.110 These findings suggest that IL-27 promotes M1 macrophage development and contributes to inflammatory damage in asthma. Interestingly, Th2-related factors such as IL-4 have been shown to enhance the expression of inflammatory cytokines, including IL-6, TNF-α, and IL-12, in macrophages challenged with Neisseria meningitidis through the MyD88 pathway.111 This indicates that IL-4 not only regulates the anti-inflammatory functions of macrophages but also promotes M1-associated cytokine expression under specific conditions. Similarly, IL-33, a Th2 cytokine, can promote the expression of M2 macrophage markers such as arginase, Ym1, and mannose receptor in M1 or M2 polarized macrophages. However, in unpolarized macrophages, IL-33 enhances the expression of M1-associated chemokine CCL3.112 This dual effect suggests that IL-33 may play a critical role in balancing inflammation and repair, though its precise mechanisms require further investigation.

M2 Polarization: Remodeling and Fibrosis in Asthma

While M1 macrophages are typically highly expressed in non-atopic asthma and are associated with severe asthma,66 M2 macrophages are considered closely related to atopic asthma and are the predominant macrophage type in this condition.63 In NM-induced lung injury mice model, macrophage polarization exhibits a time-dependent pattern: M1 macrophages are most prominent 1–3 days after NM exposure, whereas M2 macrophages peak during the later stages, particularly on day 28, contributing to lung fibrosis.113 In asthma animal models, M2 macrophages are characterized by the high expression of ARG1, chitinase-like proteins Ym1/2, Fizz1, and macrophage mannose receptor 206 (CD206).63

Arginase-1

Arginase-1 (ARG1) is one of the most representative markers of alternatively activated macrophages. It hydrolyzes L-arginine into L-ornithine and urea, facilitating nitrogen excretion.114 In macrophages, the expression of ARG1 is typically silenced but can be induced more than four orders of magnitude upon receiving Th2 cytokine signals.115 Apart from being metabolized via the ARG1 pathway, L-arginine can also be catalyzed by iNOS to produce NO and L-citrulline.116 Consequently, ARG1 in M2 macrophages competes with iNOS for the substrate L-arginine. Elevated ARG1 expression reduces the availability of L-arginine for iNOS, thereby inhibiting NO synthesis.117 NO relaxes airway smooth muscle cells, leading to bronchodilation,118 and its inhibition in asthma mouse models has been shown to induce airway hyperresponsiveness.119 As a product of the ARG1 reaction, L-ornithine is converted into polyamines, which contribute to airway hyperresponsiveness in allergen-induced asthma models. This effect can be alleviated by inhibiting ornithine decarboxylase.119 Polyamines, particularly spermine and spermidine, also stimulate smooth muscle relaxation by reducing intracellular calcium concentrations.120 Additionally, L-ornithine can be converted into proline by ornithine δ-aminotransferase.121 Proline, a critical precursor for collagen synthesis, promotes collagen deposition, leading to airway remodeling and fibrosis in asthma patients.122 These findings suggest that ARG1 regulates NO levels by competing with iNOS for L-arginine and contributes to bronchial contraction and airway remodeling through its metabolic products. As such, ARG1 plays a significant role in promoting asthma pathogenesis.

Chitinase-Like Proteins

Chitinase-like proteins (CLPs) are a group of proteins that lack traditional chitinase enzymatic activity but retain structural domains similar to chitinases. These domains allow CLPs to bind with high affinity to chitin and participate in recognizing PAMPs.123 Although the chitinase-like proteins Ym1 and Ym2 are expressed only in mice, they are functionally analogous to human galectin-10, which contributes to the formation of Charcot-Leyden crystals. Understanding the immunoregulatory roles of Ym1 and Ym2 may provide insights into asthma mechanisms in humans.124 Elevated expression of Ym1 and Ym2 has been observed in the bronchoalveolar lavage fluid (BALF) of allergic asthma mouse models, and this upregulation correlates significantly with airway inflammation.125 The expression of Ym1 and Ym2 is driven by Th2 cells and their cytokines, IL-4 and IL-13, with Ym2 particularly showing marked upregulation in the BALF of lungs affected by allergic asthma. This suggests that Ym2 may play a critical role in allergic pulmonary inflammation.126 Ym1 and Ym2 have been shown to bind to heparan sulfate and glucosamine (GlcN) oligosaccharides, such as chitobiose, chitotriose, and chitotetraose.126 Considering that heparan sulfate and GlcN oligosaccharides can bind and activate cellular responses to transforming growth factor-β (TGF-β), promoting cell proliferation and fibrotic responses,127 and that carbohydrate molecules mediate cell-matrix interactions and facilitate immune cell and fibroblast migration and localization,128 Ym1 and Ym2 may play key roles in these processes. For example, Ym1 has been identified as a chemotactic factor for eosinophils, promoting their recruitment and thus contributing to the progression of inflammation.129 These findings suggest that Ym1 and Ym2 may exert significant roles in allergic airway remodeling by interacting with carbohydrate molecules, contributing to inflammation and tissue remodeling in asthma.

Found in Inflammatory Zone 1

Fizz1, a cysteine-rich secretory protein, is predominantly expressed in immune cells, particularly macrophages.130 It is recognized as a marker of alternatively activated macrophages in mice.131 As a marker of oxidative stress, Fizz1 expression is significantly upregulated in the BALF of ovalbumin-sensitized mice.132 In airway epithelial cell brushings from mice challenged with the fungal allergen Alternaria, Fizz1 expression increased over 20-fold, an effect absents in STAT6-deficient mice, indicating that Fizz1 activation is STAT6-dependent.133 The study also revealed that Fizz1 binds to collagen-producing fibroblasts, promoting the expression of type I collagen and α-SMA in fibroblasts. This interaction contributes to peribronchial fibrosis, airway epithelial thickening, and airway remodeling.134 Similarly, Fizz1 knockout mice in a bleomycin-induced pulmonary fibrosis model exhibited impaired activation of lung fibroblasts and reduced chemotactic activity of bone marrow-derived dendritic cells, whereas Fizz1 overexpression exacerbated pulmonary fibrosis.135

Macrophage Mannose Receptor

CD206, a C-type lectin receptor expressed on the surface of macrophages, participates in immune responses by recognizing sugars such as mannose, galactose, glucose, and certain glycoproteins.136 Studies have demonstrated that CD206 expression is significantly upregulated in RAW264.7 macrophages polarized toward the M2 phenotype.137 In CD206-deficient mice, the uptake of cockroach allergens by macrophages was markedly reduced. Additionally, macrophages from CD206-/- mice exhibited significantly decreased levels of miR-511-3p and displayed an M1-like phenotype. CD206-/- mice in a cockroach allergen-induced asthma model showed exacerbated lung inflammation. Overexpression of miR-511-3p in macrophages restored an M2-like phenotype, suggesting that CD206 promotes M2 macrophage polarization by upregulating miR-511-3p.138 Further studies have shown that CD206 expression is higher in female asthma patients compared to males. This is likely due to the effects of IL-4 and IL-13, as estrogen enhances IL-4 secretion, potentially explaining the sex-related differences in CD206 expression.136 Additionally, a higher number of CD206+ M2 macrophages is associated with greater asthma severity. These macrophages are resistant to inhaled corticosteroid therapy. However, after depleting IL-10+ M2-like macrophages, the number of CD206+ macrophages decreased following corticosteroid treatment in patients with asthma. This highlights the importance of considering polarization states and the heterogeneity of macrophage populations in therapeutic strategies for asthma.136

Therapeutic Strategies Targeting Macrophage Polarization in Asthma

Macrophages perform essential immune functions in the lungs, but these functions are dysregulated in asthma.139 In recent years, the complexity and plasticity of macrophage phenotypes have been redefined, highlighting gaps in our understanding of their role in asthma pathogenesis.140 In this review, we discussed the sources of ROS and how ROS regulate macrophage polarization states. Despite the evident impact of macrophage polarization on asthma, current therapeutic strategies have largely overlooked this aspect, especially the potential of modulating macrophage polarization through ROS. Table 1 summarizes the roles of drugs targeting macrophage polarization in asthma therapy, with increasing evidence suggesting that such drugs exert therapeutic effects at least partially by modulating ROS production.

Table 1.

Therapeutic Agents Targeting Macrophage Polarization for Asthma Treatment

Therapeutic Agent Macrophage Polarization State Model Stage Effect on Asthma Treatment Mechanism of Action References
GSH Patients with asthma Pre-clinical Reduce airway hyperresponsiveness Inhibit Nrf2 and alleviate Th2 imbalance. [141]
Atorvastatin Promote M2 Polarization BALB/c mice Pre-clinical Alleviates airway inflammation Activate Nrf2, reduce ROS. [142]
N-acetylcysteine Inhibit IL-25–induced M2 macrophage polarization Human monocyte cell line Pre-clinical Inhibit IL-25–mediated mitochondrial ROS production and M2 macrophage polarization. [143]
Celastrol Promote M2 Polarization Diet-induced obesity mice Pre-clinical Alleviates airway hyperresponsiveness and inflammation Promote M1-to-M2 macrophage polarization via the PI3K/AKT pathway, reducing M1 markers (iNOS, IL-1β, TNF-α) and increasing M2 markers (Arg-1, IL-10). [144]
Calycosin Inhibit M2 Polarization BALB/C mice Pre-clinical Inhibit OVA-induced airway inflammation and remodeling. Inhibit the elevation of ARG1, IL-10, YM1, and MRC1 levels in the lung tissue of OVA-induced asthmatic mice. [145]
Quercetin Inhibit M1 Polarization C57BL/6 mice Pre-clinical Improve airway inflammation in neutrophilic asthma. Inhibit the pro-inflammatory M1 phenotype of macrophages and reduce inflammatory cytokine and MDA levels. [146]
AUDA Promote M2 Polarization Obese asthmatic mice Pre-clinical Alleviates airway hyperresponsiveness and inflammation Reduce the elevation of IL-1β, IL-6, and TNF-α in RAW264.7 macrophages stimulated by LPS. [147]
Ruxolitinib Suppress M2 macrophage polarization C57BL/6 mice Pre-clinical Inhibit OVA-induced allergic asthma Inhibit mitochondrial ROS-dependent STAT6 activation, thereby weakening DMM-induced M2 macrophage activation. [148]

Abbreviations: AKT, Protein kinase B; Arg-1, Arginase-1; AUDA, 12-(3-Adamantan-1-yl-ureido)-dodecanoic acid; DMM, Dimethyl malonate; GSH, Reduced glutathione; iNOS, Inducible nitric oxide synthase; LPS, Lipopolysaccharide; MDA, Malondialdehyde; MRC1, Mannose receptor C type 1; Nrf2, Nuclear factor erythroid 2 related factor 2; PI3K, Phosphatidylinositol 3 kinase; STAT6, Signal transducer and activator of transcription 6; YM1, Chitinase like protein 3.

Studies have shown that AMs in children with poorly controlled asthma exhibit significantly reduced phagocytic capacity.149 Fitzpatrick et al reported elevated oxidative stress markers, including glutathione disulfide (GSSG), oxidative DNA byproducts, and inflammatory levels, in the ELF of children with severe asthma. This was accompanied by diminished phagocytic ability, which could be ameliorated by the antioxidant reduced glutathione (GSH). GSH restored phagocytic function, reduced airway hyperresponsiveness, and mitigated Th2 imbalance.141 Since M2 macrophages generally have lower phagocytic capacity than M1 macrophages,139 ROS may influence asthma progression by modulating macrophage polarization states. Soluble epoxide hydrolase (sEH), which converts antioxidant epoxyeicosatrienoic acids into diols, contributes to oxidative stress. Atorvastatin has been reported to decrease MDA levels while elevating SOD activity in mice, concurrently suppressing M1 macrophage polarization and favoring the M2 phenotype.150 In an OVA murine asthma model, atorvastatin also activated the Nrf2 pathway and mitigated airway inflammation.142 Given that Nrf2 activation lowers ROS production, restrains M1 polarization, and alleviates pulmonary inflammation,151 it is plausible that atorvastatin exerts anti-asthmatic effects by attenuating ROS and thereby limiting M1-driven responses. Nonetheless, dedicated studies are still required to verify atorvastatin’s macrophage-modulating actions in asthma. Beyond statins, IL-25—a prototypical Th2 cytokine—has been shown to enhance mitochondrial respiratory-chain complex activity, elevate ROS, activate AMPK, and ultimately drive M2 polarization of monocyte-derived macrophages.143 Notably, N-acetyl-l-cysteine suppresses IL-25-induced ROS generation and consequently blocks this M2-skewing effect.143 Collectively, these findings underscore the therapeutic potential of targeting mitochondrial ROS and redox signaling to control macrophage polarization and airway inflammation in asthma.

Some plant extracts with strong antioxidant properties have shown potential in regulating macrophage polarization and improving asthma outcomes. Celastrol, derived from the roots of Tripterygium wilfordii, enhances the expression of heme oxygenase-1 (HO-1) to exert antioxidant effects.152 Celastrol suppresses M1 polarization of AMs, promotes M2 polarization, and reduces airway inflammatory cell infiltration and goblet cell hyperplasia in diet-induced obesity asthma mice.144 Calycosin, a flavonoid extracted from Astragalus membranaceus, inhibits mtROS production by increasing levels of SOD and GSH and reducing malondialdehyde levels.153 Calycosin has been reported to suppress elevated levels of ARG1, IL-10, CLP, and Mrc1 in the lung tissues of OVA-induced asthma mouse models, thereby mitigating airway inflammation and remodeling.145 The natural flavonoid quercetin inhibits NOX2, thereby reducing LPS-induced intracellular ROS levels.154 Quercetin also suppresses the pro-inflammatory M1 phenotype of macrophages, alleviating airway inflammation in LPS/OVA-induced asthma mouse models.146 These findings highlight the therapeutic potential of targeting macrophage polarization and ROS regulation in asthma management. Further exploration of these strategies may provide more effective treatments for different asthma phenotypes.

Beyond plant‐derived extracts, studies have shown that sEH inhibitors, such as 12-(3-Adamantan-1-yl-ureido)-dodecanoic acid (AUDA), can shift macrophages from an M1 to an M2 phenotype, reducing airway inflammation and hyperresponsiveness in obese asthma mouse models.147 Dimethyl malonate, an inhibitor of succinate dehydrogenase, suppresses dehydrogenase activity and, through mitochondrial ROS–dependent STAT6 activation, up-regulates M2-signature genes Arg1, Ym1, and Mrc1, thereby aggravating OVA-induced allergic asthma in vivo. Scavenging mitochondrial ROS with mitoTEMPO or blocking STAT6 activation with ruxolitinib mitigates the pro-M2 effect of DMM.148 Collectively, these data identify dehydrogenase as a mitochondrial ROS gated brake on M2 polarization and underscore its potential as a therapeutic target in M2-driven asthma, although further work is needed to translate dehydrogenase centered interventions into clinical practice.

Conclusion and Future Perspectives

An increasing body of research highlights the critical role of macrophages in the progression of asthma, with ROS acting as key signaling molecules that promote macrophage activation. ROS are derived from both endogenous and exogenous sources. Endogenous ROS primarily include intracellular oxidase systems (mainly NOX, XO, and peroxisomes) and mtROS, while exogenous ROS are predominantly generated by air pollutants. ROS regulate macrophage polarization states through intracellular signaling pathways. Although ROS have been shown to influence macrophage polarization, the specific inclination of ROS toward particular polarization states remains unclear. Given that ROS from different sources exhibit distinct roles across various stages of M2 polarization, future research should focus on understanding the regulatory effects of ROS from diverse origins at different macrophage developmental stages to clarify the directional influence of ROS on polarization.

In summary, ROS and macrophage polarization constitute a central pathogenic axis that drives airway inflammation and remodeling in asthma. Broad-spectrum antioxidants can attenuate airway hyperresponsiveness, but their lack of cellular specificity prevents precise control of the M1 to M2 transition. Future work should map the pathways of ROS production and clearance within macrophage subsets and design redox-modulating agents, for example nanoparticle-based delivery systems, that selectively target macrophages and steer their polarization. Combining ROS modulation with metabolic reprogramming, cytokine blockade, or signaling pathway inhibition is likely to generate multitarget strategies that suppress inflammation while bypassing the limitations of single antioxidant interventions. Research should also move beyond indiscriminate ROS scavenging toward preservation of redox homeostasis, which restrains excessive inflammation without sacrificing macrophage functional diversity. Biomarkers that dynamically report macrophage polarization and ROS levels will facilitate patient stratification and assessment of therapeutic efficacy, accelerating clinical translation of lead compounds. A more detailed mechanistic and temporal portrait of macrophage-derived ROS should deliver refined, multifaceted therapeutic options and ultimately improve outcomes in refractory asthma.

Funding Statement

This work was funded by the Scientific Research Project of Higher Education Institutions in Anhui Province (Natural Science) (No. 2022AH050442 and No. 2023AH050723) and Anhui University of Chinese Medicine 2019-2020 Talent Support Program (No. DT2000000458).

Abbreviations

AMs, alveolar macrophages; ROS, reactive oxygen species; NADPH, nicotinamide adenine dinucleotide phosphate; NOX, NADPH oxidase; PKC, protein kinase C; TLRs, Toll-like receptors; GEFs, guanine nucleotide exchange factors; mtROS, mitochondrial ROS; MAPK, mitogen activated protein kinase; NFAT5, nuclear factor of activated T cells 5; FAD, flavin adenine dinucleotide; TCA, tricarboxylic acid; FADH2, reduced flavin adenine dinucleotide; FMN, flavin mononucleotide; CoQ, coenzyme Q; SOD, superoxide dismutase; LPS, lipopolysaccharide; HIF-1α, hypoxia-inducible factor 1-alpha; O₃, ozone; PM, particulate matter; ELF, epithelial lining fluid; PM10, PM with a diameter of less than 10 micrometers; PAHs, polycyclic aromatic hydrocarbons; JAK, Janus kinase; STAT, signal transducer and activator of transcription; TRIF, TIR-domain-containing adapter-inducing interferon-β; iNOS, inducible nitric oxide synthase; NF-κB, nuclear factor kappa-light-chain-enhancer of activated B cells; Trx, thioredoxin; ASK1, apoptosis signal-regulating kinase 1; CHK2, checkpoint kinase 2; PKM2, pyruvate kinase M2; DAMPs, damage-associated molecular patterns; PAMPs, pathogen-associated molecular patterns; NLRP3, NOD-like receptor family pyrin domain containing 3; ARG1, arginase-1; SYNCRIP, synaptotagmin-binding cytoplasmic RNA-interacting protein; Ym1, chitinase 3-like protein 3; Fizz1, found in inflammatory zone 1; Mrc1, mannose receptor c-type 1; HDM, house dust mite; FDE, farm dust extract; NO, nitric oxide; GR, glucocorticoid receptor; CD206, mannose receptor 206; ARG1, Arginase-1; CLPs, Chitinase-like proteins; BALF, bronchoalveolar lavage fluid; GlcN, glucosamine; TGF-β, transforming growth factor-β; sEH, Soluble epoxide hydrolase; AUDA, 12-(3-Adamantan-1-yl-ureido)-dodecanoic acid; HO-1, heme oxygenase-1.

Author Contributions

All authors made a significant contribution to the work reported, whether that is in the conception, study design, execution, acquisition of data, analysis and interpretation, or in all these areas; took part in drafting, revising or critically reviewing the article; gave final approval of the version to be published; have agreed on the journal to which the article has been submitted; and agree to be accountable for all aspects of the work.

Disclosure

The authors declare that they have no competing interests.

References

  • 1.Porsbjerg C, Melén E, Lehtimäki L, Shaw D. Asthma. Lancet. 2023;401:858–873. doi: 10.1016/S0140-6736(22)02125-0 [DOI] [PubMed] [Google Scholar]
  • 2.Levy ML, Bacharier LB, Bateman E, et al. Key recommendations for primary care from the 2022 global initiative for asthma (GINA) update. NPJ Prim Care Respir Med. 2023;33:7. doi: 10.1038/s41533-023-00330-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Xie C, Yang J, Gul A, et al. Immunologic aspects of asthma: from molecular mechanisms to disease pathophysiology and clinical translation. Front Immunol. 2024;15:1478624. doi: 10.3389/fimmu.2024.1478624 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Do DC, Zhao Y, Gao P. Cockroach allergen exposure and risk of asthma. Allergy. 2016;71:463–474. doi: 10.1111/all.12827 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Delfino RJ, Staimer N, Tjoa T, Gillen DL, Schauer JJ, Shafer MM. Airway inflammation and oxidative potential of air pollutant particles in a pediatric asthma panel. J Expo Sci Environ Epidemiol. 2013;23:466–473. doi: 10.1038/jes.2013.25 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Cai Y, Sugimoto C, Arainga M, Alvarez X, Didier ES, Kuroda MJ. In vivo characterization of alveolar and interstitial lung macrophages in rhesus macaques: implications for understanding lung disease in humans. J Immunol. 2014;192:2821–2829. doi: 10.4049/jimmunol.1302269 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Hou F, Xiao K, Tang L, Xie L. Diversity of macrophages in lung homeostasis and diseases. Front Immunol. 2021;12:753940. doi: 10.3389/fimmu.2021.753940 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Zhong Y, Huang T, Huang J, et al. The HDAC10 instructs macrophage M2 program via deacetylation of STAT3 and promotes allergic airway inflammation. Theranostics. 2023;13:3568–3581. doi: 10.7150/thno.82535 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Wang Y, Le Y, Wu J, et al. Inhibition of xanthine oxidase by allopurinol suppresses HMGB1 secretion and ameliorates experimental asthma. Redox Biol. 2024;70:103021. doi: 10.1016/j.redox.2023.103021 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Wood LG, Wark PA, Garg ML. Antioxidant and anti-inflammatory effects of resveratrol in airway disease. Antioxid Redox Signal. 2010;13:1535–1548. doi: 10.1089/ars.2009.3064 [DOI] [PubMed] [Google Scholar]
  • 11.Liu J, Wei Y, Jia W, et al. Chenodeoxycholic acid suppresses AML progression through promoting lipid peroxidation via ROS/p38 MAPK/DGAT1 pathway and inhibiting M2 macrophage polarization. Redox Biol. 2022;56:102452. doi: 10.1016/j.redox.2022.102452 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Hahner F, Moll F, Schröder K. NADPH oxidases in the differentiation of endothelial cells. Cardiovasc Res. 2020;116:262–268. doi: 10.1093/cvr/cvz213 [DOI] [PubMed] [Google Scholar]
  • 13.Bedard K, Krause K-H. The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology. Physiol Rev. 2007;87:245–313. doi: 10.1152/physrev.00044.2005 [DOI] [PubMed] [Google Scholar]
  • 14.Bode K, Hauri-Hohl M, Jaquet V, Weyd H. Unlocking the power of NOX2: a comprehensive review on its role in immune regulation. Redox Biol. 2023;64:102795. doi: 10.1016/j.redox.2023.102795 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Paclet MH, Laurans S, Dupré-Crochet S. Regulation of neutrophil NADPH oxidase, NOX2: a crucial effector in neutrophil phenotype and function. Front Cell Dev Biol. 2022;10:945749. doi: 10.3389/fcell.2022.945749 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.El-Benna J, Dang PM, Gougerot-Pocidalo MA, Marie JC, Braut-Boucher F. p47phox, the phagocyte NADPH oxidase/NOX2 organizer: structure, phosphorylation and implication in diseases. Exp Mol Med. 2009;41:217–225. doi: 10.3858/emm.2009.41.4.058 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Park JW, Babior BM. Activation of the leukocyte NADPH oxidase subunit p47phox by protein kinase C. A phosphorylation-dependent change in the conformation of the C-terminal end of p47phox. Biochemistry. 1997;36:7474–7480. [DOI] [PubMed] [Google Scholar]
  • 18.El-Benna J, Hurtado-Nedelec M, Marzaioli V, Marie JC, Gougerot-Pocidalo MA, Dang PM. Priming of the neutrophil respiratory burst: role in host defense and inflammation. Immunol Rev. 2016;273:180–193. [DOI] [PubMed] [Google Scholar]
  • 19.Hordijk PL. Regulation of NADPH oxidases: the role of Rac proteins. Circ Res. 2006;98:453–462. doi: 10.1161/01.RES.0000204727.46710.5e [DOI] [PubMed] [Google Scholar]
  • 20.Sareila O, Kelkka T, Pizzolla A, Hultqvist M, Holmdahl R. NOX2 complex-derived ROS as immune regulators. Antioxid Redox Signal. 2011;15:2197–2208. doi: 10.1089/ars.2010.3635 [DOI] [PubMed] [Google Scholar]
  • 21.Zhang X, Wang Q, Wu J, Wang J, Shi Y, Liu M. Crystal structure of human lysyl oxidase-like 2 (hLOXL2) in a precursor state. Proc Natl Acad Sci U S A. 2018;115:3828–3833. doi: 10.1073/pnas.1720859115 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Schmidt HM, Kelley EE, Straub AC. The impact of xanthine oxidase (XO) on hemolytic diseases. Redox Biol. 2019;21:101072. doi: 10.1016/j.redox.2018.101072 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Dai Y, Cao Y, Zhang Z, Vallurupalli S, Mehta JL. Xanthine oxidase induces foam cell formation through LOX-1 and NLRP3 activation. Cardiovasc Drugs Ther. 2017;31:19–27. doi: 10.1007/s10557-016-6706-x [DOI] [PubMed] [Google Scholar]
  • 24.Sarniak A, Lipińska J, Tytman K, Lipińska S. Endogenous mechanisms of reactive oxygen species (ROS) generation. Postepy Hig Med Dosw. 2016;70:1150–1165. doi: 10.5604/17322693.1224259 [DOI] [PubMed] [Google Scholar]
  • 25.Kim N-H, Choi S, Han E-J, et al. The xanthine oxidase–NFAT5 pathway regulates macrophage activation and TLR-induced inflammatory arthritis. Eur J Immunol. 2014;44:2721–2736. doi: 10.1002/eji.201343669 [DOI] [PubMed] [Google Scholar]
  • 26.Ives A, Nomura J, Martinon F, et al. Xanthine oxidoreductase regulates macrophage IL1β secretion upon NLRP3 inflammasome activation. Nat Commun. 2015;6:6555. doi: 10.1038/ncomms7555 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Huang L, Liu Y, Wang X, et al. Peroxisome-mediated reactive oxygen species signals modulate programmed cell death in plants. Int J Mol Sci. 2022;23:10087. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Shai N, Yifrach E, van Roermund CWT, et al. Systematic mapping of contact sites reveals tethers and a function for the peroxisome-mitochondria contact. Nat Commun. 2018;9:1761. doi: 10.1038/s41467-018-03957-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Nakamura H, Fang J, Maeda H. Protective role of d-amino acid oxidase against staphylococcus aureus infection. Infect Immun. 2012;80:1546–1553. doi: 10.1128/IAI.06214-11 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Liu Y, Chen W, Li C, et al. DsbA-L interacting with catalase in peroxisome improves tubular oxidative damage in diabetic nephropathy. Redox Biol. 2023;66:102855. doi: 10.1016/j.redox.2023.102855 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Baker A, Graham IA, Holdsworth M, Smith SM, Theodoulou FL. Chewing the fat: beta-oxidation in signalling and development. Trends Plant Sci. 2006;11:124–132. doi: 10.1016/j.tplants.2006.01.005 [DOI] [PubMed] [Google Scholar]
  • 32.Yu L, Fan J, Xu C. Peroxisomal fatty acid β-oxidation negatively impacts plant survival under salt stress. Plant Signal Behav. 2019;14:1561121. doi: 10.1080/15592324.2018.1561121 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Dan Dunn J, Alvarez LA, Zhang X, Soldati T. Reactive oxygen species and mitochondria: a nexus of cellular homeostasis. Redox Biol. 2015;6:472–485. doi: 10.1016/j.redox.2015.09.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.West AP, Shadel GS, Ghosh S. Mitochondria in innate immune responses. Nat Rev Immunol. 2011;11:389–402. doi: 10.1038/nri2975 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Murphy MP. How mitochondria produce reactive oxygen species. Biochem J. 2009;417:1–13. doi: 10.1042/BJ20081386 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Grba DN, Chung I, Bridges HR, Agip AA, Hirst J. Investigation of hydrated channels and proton pathways in a high-resolution cryo-EM structure of mammalian complex I. Sci Adv. 2023;9:eadi1359. doi: 10.1126/sciadv.adi1359 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Hirst J, King MS, Pryde KR. The production of reactive oxygen species by complex I. Biochem Soc Trans. 2008;36:976–980. doi: 10.1042/BST0360976 [DOI] [PubMed] [Google Scholar]
  • 38.Iwata S, Lee JW, Okada K, et al. Complete structure of the 11-subunit bovine mitochondrial cytochrome bc1 complex. Science. 1998;281:64–71. doi: 10.1126/science.281.5373.64 [DOI] [PubMed] [Google Scholar]
  • 39.Ghanian Z, Konduri GG, Audi SH, Camara AKS, Ranji M. Quantitative optical measurement of mitochondrial superoxide dynamics in pulmonary artery endothelial cells. J Innov Opt Health Sci. 2018;11. doi: 10.1142/S1793545817500183 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Kudin AP, Bimpong-Buta NY, Vielhaber S, Elger CE, Kunz WS. Characterization of superoxide-producing sites in isolated brain mitochondria. J Biol Chem. 2004;279:4127–4135. doi: 10.1074/jbc.M310341200 [DOI] [PubMed] [Google Scholar]
  • 41.Brand MD. The sites and topology of mitochondrial superoxide production. Exp Gerontol. 2010;45:466–472. doi: 10.1016/j.exger.2010.01.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Han D, Antunes F, Canali R, Rettori D, Cadenas E. Voltage-dependent anion channels control the release of the superoxide anion from mitochondria to cytosol. J Biol Chem. 2003;278:5557–5563. doi: 10.1074/jbc.M210269200 [DOI] [PubMed] [Google Scholar]
  • 43.Hadrava Vanova K, Kraus M, Neuzil J, Rohlena J. Mitochondrial complex II and reactive oxygen species in disease and therapy. Redox Rep. 2020;25:26–32. doi: 10.1080/13510002.2020.1752002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Reczek CR, Chandel NS. ROS-dependent signal transduction. Curr Opin Cell Biol. 2015;33:8–13. doi: 10.1016/j.ceb.2014.09.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.West AP, Brodsky IE, Rahner C, et al. TLR signalling augments macrophage bactericidal activity through mitochondrial ROS. Nature. 2011;472:476–480. doi: 10.1038/nature09973 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Mills EL, Kelly B, Logan A, et al. Succinate dehydrogenase supports metabolic repurposing of mitochondria to drive inflammatory macrophages. Cell. 2016;167:457–470.e413. doi: 10.1016/j.cell.2016.08.064 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Fuhrmann DC, Wittig I, Brüne B. TMEM126B deficiency reduces mitochondrial SDH oxidation by LPS, attenuating HIF-1α stabilization and IL-1β expression. Redox Biol. 2019;20:204–216. doi: 10.1016/j.redox.2018.10.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Gu L, Larson Casey JL, Andrabi SA, et al. Mitochondrial calcium uniporter regulates PGC-1α expression to mediate metabolic reprogramming in pulmonary fibrosis. Redox Biol. 2019;26:101307. doi: 10.1016/j.redox.2019.101307 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Jacquemin B, Sunyer J, Forsberg B, et al. Home outdoor NO2 and new onset of self-reported asthma in adults. Epidemiology. 2009;20:119–126. doi: 10.1097/EDE.0b013e3181886e76 [DOI] [PubMed] [Google Scholar]
  • 50.Liccardi G, D’Amato G, D’Amato M, Holgate S. Environmental risk factors and allergic bronchial asthma. Clin Exp Allergy. 2005;35:1113–1124. doi: 10.1111/j.1365-2222.2005.02328.x [DOI] [PubMed] [Google Scholar]
  • 51.Tapak M, Sadeghi S, Ghazanfari T, Mosaffa N. Chemical exposure and alveolar macrophages responses: ‘the role of pulmonary defense mechanism in inhalation injuries’. BMJ Open Respir Res. 2023;10:e001589. doi: 10.1136/bmjresp-2022-001589 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Kelly FJ, Mudway I, Krishna MT, Holgate ST. The free radical basis of air pollution: focus on ozone. Respir Med. 1995;89:647–656. doi: 10.1016/0954-6111(95)90131-0 [DOI] [PubMed] [Google Scholar]
  • 53.Michaudel C, Mackowiak C, Maillet I, et al. Ozone exposure induces respiratory barrier biphasic injury and inflammation controlled by IL-33. J Allergy Clin Immunol. 2018;142:942–958. doi: 10.1016/j.jaci.2017.11.044 [DOI] [PubMed] [Google Scholar]
  • 54.Qu J, Li Y, Zhong W, Gao P, Hu C. Recent developments in the role of reactive oxygen species in allergic asthma. J Thorac Dis. 2017;9:E32–e43. doi: 10.21037/jtd.2017.01.05 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Cergan R, Berghi O, Dumitru M, et al. Interleukin 8 molecular interplay in allergic rhinitis and chronic rhinosinusitis with nasal polyps: a scoping review. Life. 2025;15:469. doi: 10.3390/life15030469 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Cambier S, Gouwy M, Proost P. The chemokines CXCL8 and CXCL12: molecular and functional properties, role in disease and efforts towards pharmacological intervention. Cell Mol Immunol. 2023;20:217–251. doi: 10.1038/s41423-023-00974-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Sarnat JA, Schwartz J, Suh HH. Fine particulate air pollution and mortality in 20 U.S. cities. N Engl J Med. 2001;344:1253–1254. [DOI] [PubMed] [Google Scholar]
  • 58.Strak M, Janssen N, Beelen R, et al. Long-term exposure to particulate matter, NO(2) and the oxidative potential of particulates and diabetes prevalence in a large national health survey. Environ Int. 2017;108:228–236. doi: 10.1016/j.envint.2017.08.017 [DOI] [PubMed] [Google Scholar]
  • 59.Michaeloudes C, Abubakar-Waziri H, Lakhdar R, et al. Molecular mechanisms of oxidative stress in asthma. Mol Aspects Med. 2022;85:101026. doi: 10.1016/j.mam.2021.101026 [DOI] [PubMed] [Google Scholar]
  • 60.Li N, Sioutas C, Cho A, et al. Ultrafine particulate pollutants induce oxidative stress and mitochondrial damage. Environ Health Perspect. 2003;111:455–460. doi: 10.1289/ehp.6000 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Penning TM, Burczynski ME, Hung CF, McCoull KD, Palackal NT, Tsuruda LS. Dihydrodiol dehydrogenases and polycyclic aromatic hydrocarbon activation: generation of reactive and redox active o-quinones. Chem Res Toxicol. 1999;12:1–18. doi: 10.1021/tx980143n [DOI] [PubMed] [Google Scholar]
  • 62.Shafer MM, Perkins DA, Antkiewicz DS, Stone EA, Quraishi TA, Schauer JJ. Reactive oxygen species activity and chemical speciation of size-fractionated atmospheric particulate matter from Lahore, Pakistan: an important role for transition metals. J Environ Monit. 2010;12:704–715. doi: 10.1039/B915008K [DOI] [PubMed] [Google Scholar]
  • 63.Luo M, Zhao F, Cheng H, Su M, Wang Y. Macrophage polarization: an important role in inflammatory diseases. Front Immunol. 2024;15:1352946. doi: 10.3389/fimmu.2024.1352946 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Varricchi G, Brightling CE, Grainge C, Lambrecht BN, Chanez P. Airway remodelling in asthma and the epithelium: on the edge of a new era. Eur Respir J. 2024;63. doi: 10.1183/13993003.01619-2023 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Sharma N, Akkoyunlu M, Rabin RL. Macrophages-common culprit in obesity and asthma. Allergy. 2018;73:1196–1205. doi: 10.1111/all.13369 [DOI] [PubMed] [Google Scholar]
  • 66.Saradna A, Do DC, Kumar S, Fu QL, Gao P. Macrophage polarization and allergic asthma. Transl Res. 2018;191:1–14. doi: 10.1016/j.trsl.2017.09.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Formentini L, Santacatterina F, Núñez de Arenas C, et al. Mitochondrial ROS production protects the intestine from inflammation through functional M2 macrophage polarization. Cell Rep. 2017;19:1202–1213. doi: 10.1016/j.celrep.2017.04.036 [DOI] [PubMed] [Google Scholar]
  • 68.Li C, Deng C, Wang S, et al. A novel role for the ROS-ATM-Chk2 axis mediated metabolic and cell cycle reprogramming in the M1 macrophage polarization. Redox Biol. 2024;70:103059. doi: 10.1016/j.redox.2024.103059 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Horvath CM. The Jak-STAT pathway stimulated by interferon γ. Science’s STKE. 2004;2004:tr8–tr8. doi: 10.1126/stke.2602004tr8 [DOI] [PubMed] [Google Scholar]
  • 70.Guo L, Chen S, Liu Q, et al. Glutaredoxin 1 regulates macrophage polarization through mediating glutathionylation of STAT1. Thoracic Cancer. 2020;11:2966–2974. doi: 10.1111/1759-7714.13647 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Sica A, Erreni M, Allavena P, Porta C. Macrophage polarization in pathology. Cell Mol Life Sci. 2015;72:4111–4126. doi: 10.1007/s00018-015-1995-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Lauterbach MA, Hanke JE, Serefidou M, et al. Toll-like receptor signaling rewires macrophage metabolism and promotes histone acetylation via ATP-citrate lyase. Immunity. 2019;51:997–1011.e1017. doi: 10.1016/j.immuni.2019.11.009 [DOI] [PubMed] [Google Scholar]
  • 73.Hu D, Wang Y, You Z, Lu Y, Liang C. lnc-MRGPRF-6:1 promotes M1 polarization of macrophage and inflammatory response through the TLR4-MyD88-MAPK pathway. Mediators Inflamm. 2022;2022:6979117. doi: 10.1155/2022/6979117 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Li Y, Zhang L, Ren P, et al. Qing-Xue-Xiao-Zhi formula attenuates atherosclerosis by inhibiting macrophage lipid accumulation and inflammatory response via TLR4/MyD88/NF-κB pathway regulation. Phytomedicine. 2021;93:153812. doi: 10.1016/j.phymed.2021.153812 [DOI] [PubMed] [Google Scholar]
  • 75.Gorina R, Font-Nieves M, Márquez-Kisinousky L, Santalucia T, Planas AM. Astrocyte TLR4 activation induces a proinflammatory environment through the interplay between MyD88-dependent NFκB signaling, MAPK, and Jak1/Stat1 pathways. Glia. 2011;59:242–255. doi: 10.1002/glia.21094 [DOI] [PubMed] [Google Scholar]
  • 76.Garibotto G, Carta A, Picciotto D, Viazzi F, Verzola D. Toll-like receptor-4 signaling mediates inflammation and tissue injury in diabetic nephropathy. J Nephrol. 2017;30:719–727. doi: 10.1007/s40620-017-0432-8 [DOI] [PubMed] [Google Scholar]
  • 77.La O, Bowie AG. The family of five: TIR-domain-containing adaptors in Toll-like receptor signalling. Nat Rev Immunol. 2007;7:353–364. doi: 10.1038/nri2079 [DOI] [PubMed] [Google Scholar]
  • 78.Krausgruber T, Blazek K, Smallie T, et al. IRF5 promotes inflammatory macrophage polarization and TH1-TH17 responses. Nat Immunol. 2011;12:231–238. doi: 10.1038/ni.1990 [DOI] [PubMed] [Google Scholar]
  • 79.Yamamoto M, Sato S, Hemmi H, et al. Role of adaptor TRIF in the MyD88-independent toll-like receptor signaling pathway. Science. 2003;301:640–643. doi: 10.1126/science.1087262 [DOI] [PubMed] [Google Scholar]
  • 80.Pawate S, Shen Q, Fan F, Bhat NR. Redox regulation of glial inflammatory response to lipopolysaccharide and interferongamma. J Neurosci Res. 2004;77:540–551. doi: 10.1002/jnr.20180 [DOI] [PubMed] [Google Scholar]
  • 81.Kohchi C, Inagawa H, Nishizawa T, Soma G. ROS and innate immunity. Anticancer Res. 2009;29:817–821. [PubMed] [Google Scholar]
  • 82.Liu H, Ichijo H, Nishitoh H, Kyriakis JM. Activation of apoptosis signal-regulating kinase 1 (ASK1) by tumor necrosis factor receptor-associated factor 2 requires prior dissociation of the ASK1 inhibitor thioredoxin. Mol Cell Biol. 2000;20:2198–2208. doi: 10.1128/MCB.20.6.2198-2208.2000 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Zhang J, Liu X, Wan C, et al. NLRP3 inflammasome mediates M1 macrophage polarization and IL-1β production in inflammatory root resorption. J Clin Periodontol. 2020;47:451–460. doi: 10.1111/jcpe.13258 [DOI] [PubMed] [Google Scholar]
  • 84.Wiegman CH, Li F, Ryffel B, Togbe D, Chung KF. Oxidative stress in ozone-induced chronic lung inflammation and emphysema: a facet of chronic obstructive pulmonary disease. Front Immunol. 2020;11:1957. doi: 10.3389/fimmu.2020.01957 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Li P, Li S, Wang L, et al. Mitochondrial dysfunction in hearing loss: oxidative stress, autophagy and NLRP3 inflammasome. Front Cell Dev Biol. 2023;11:1119773. doi: 10.3389/fcell.2023.1119773 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Dashdorj A, Jyothi KR, Lim S, et al. Mitochondria-targeted antioxidant MitoQ ameliorates experimental mouse colitis by suppressing NLRP3 inflammasome-mediated inflammatory cytokines. BMC Med. 2013;11:178. doi: 10.1186/1741-7015-11-178 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.van Bruggen R, Köker MY, Jansen M, et al. Human NLRP3 inflammasome activation is Nox1-4 independent. Blood. 2010;115:5398–5400. doi: 10.1182/blood-2009-10-250803 [DOI] [PubMed] [Google Scholar]
  • 88.Murata Y, Shimamura T, Hamuro J. The polarization of T(h)1/T(h)2 balance is dependent on the intracellular thiol redox status of macrophages due to the distinctive cytokine production. Int Immunol. 2002;14:201–212. doi: 10.1093/intimm/14.2.201 [DOI] [PubMed] [Google Scholar]
  • 89.Kuchler L, Giegerich AK, Sha LK, et al. SYNCRIP-dependent Nox2 mRNA destabilization impairs ROS formation in M2-polarized macrophages. Antioxid Redox Signal. 2014;21:2483–2497. doi: 10.1089/ars.2013.5760 [DOI] [PubMed] [Google Scholar]
  • 90.Padgett LE, Burg AR, Lei W, Tse HM. Loss of NADPH oxidase-derived superoxide skews macrophage phenotypes to delay type 1 diabetes. Diabetes. 2015;64:937–946. doi: 10.2337/db14-0929 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Yi L, Liu Q, Orandle MS, et al. p47(phox) directs murine macrophage cell fate decisions. Am J Pathol. 2012;180:1049–1058. doi: 10.1016/j.ajpath.2011.11.019 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Choi SH, Aid S, Kim HW, Jackson SH, Bosetti F. Inhibition of NADPH oxidase promotes alternative and anti-inflammatory microglial activation during neuroinflammation. J Neurochem. 2012;120:292–301. doi: 10.1111/j.1471-4159.2011.07572.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Zhang Y, Choksi S, Chen K, Pobezinskaya Y, Linnoila I, Liu ZG. ROS play a critical role in the differentiation of alternatively activated macrophages and the occurrence of tumor-associated macrophages. Cell Res. 2013;23:898–914. doi: 10.1038/cr.2013.75 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Shan M, Qin J, Jin F, et al. Autophagy suppresses isoprenaline-induced M2 macrophage polarization via the ROS/ERK and mTOR signaling pathway. Free Radic Biol Med. 2017;110:432–443. doi: 10.1016/j.freeradbiomed.2017.05.021 [DOI] [PubMed] [Google Scholar]
  • 95.Sica A, Bronte V. Altered macrophage differentiation and immune dysfunction in tumor development. J Clin Invest. 2007;117:1155–1166. doi: 10.1172/JCI31422 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.He C, Ryan AJ, Murthy S, Carter AB. Accelerated development of pulmonary fibrosis via Cu, Zn-superoxide dismutase-induced alternative activation of macrophages. J Biol Chem. 2013;288:20745–20757. doi: 10.1074/jbc.M112.410720 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Ji WJ, Ma YQ, Zhou X, et al. Temporal and spatial characterization of mononuclear phagocytes in circulating, lung alveolar and interstitial compartments in a mouse model of bleomycin-induced pulmonary injury. J Immunol Methods. 2014;403:7–16. doi: 10.1016/j.jim.2013.11.012 [DOI] [PubMed] [Google Scholar]
  • 98.Ross EA, Devitt A, Johnson JR. Macrophages: The good, the bad, and the gluttony. Front Immunol. 2021;12:708186. doi: 10.3389/fimmu.2021.708186 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Robbe P, Draijer C, Borg TR, et al. Distinct macrophage phenotypes in allergic and nonallergic lung inflammation. Am J Physiol Lung Cell Mol Physiol. 2015;308:L358–367. doi: 10.1152/ajplung.00341.2014 [DOI] [PubMed] [Google Scholar]
  • 100.Veremeyko T, Siddiqui S, Sotnikov I, Yung A, Ponomarev ED. IL-4/IL-13-dependent and independent expression of miR-124 and its contribution to M2 phenotype of monocytic cells in normal conditions and during allergic inflammation. PLoS One. 2013;8:e81774. doi: 10.1371/journal.pone.0081774 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Blasi F, Johnston SL. The role of antibiotics in asthma. Int J Antimicrob Agents. 2007;29:485–493. doi: 10.1016/j.ijantimicag.2006.11.029 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Khanduja KL, Kaushik G, Khanduja S, Pathak CM, Laldinpuii J, Behera D. Corticosteroids affect nitric oxide generation, total free radicals production, and nitric oxide synthase activity in monocytes of asthmatic patients. Mol Cell Biochem. 2011;346:31–37. doi: 10.1007/s11010-010-0588-1 [DOI] [PubMed] [Google Scholar]
  • 103.Naura AS, Zerfaoui M, Kim H, et al. Requirement for inducible nitric oxide synthase in chronic allergen exposure-induced pulmonary fibrosis but not inflammation. J Immunol. 2010;185:3076–3085. doi: 10.4049/jimmunol.0904214 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Tang C, Rolland JM, Li X, Ward C, Bish R, Walters EH. Alveolar macrophages from atopic asthmatics, but not atopic nonasthmatics, enhance interleukin-5 production by CD4+ T cells. Am J Respir Crit Care Med. 1998;157:1120–1126. doi: 10.1164/ajrccm.157.4.9706118 [DOI] [PubMed] [Google Scholar]
  • 105.Weltman JK, Karim AS. Interleukin-5: a proeosinophil cytokine mediator of inflammation in asthma and a target for antisense therapy. Allergy Asthma Proc. 1998;19:257–261. doi: 10.2500/108854198778557782 [DOI] [PubMed] [Google Scholar]
  • 106.Gallelli L, Falcone D, Pelaia G, et al. Interleukin-6 receptor superantagonist Sant7 inhibits TGF-beta-induced proliferation of human lung fibroblasts. Cell Prolif. 2008;41:393–407. doi: 10.1111/j.1365-2184.2008.00538.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Kim YK, Oh SY, Jeon SG, et al. Airway exposure levels of lipopolysaccharide determine type 1 versus type 2 experimental asthma. J Immunol. 2007;178:5375–5382. doi: 10.4049/jimmunol.178.8.5375 [DOI] [PubMed] [Google Scholar]
  • 108.Li JJ, Wang W, Baines KJ, et al. IL-27/IFN-γ induce MyD88-dependent steroid-resistant airway hyperresponsiveness by inhibiting glucocorticoid signaling in macrophages. J Immunol. 2010;185:4401–4409. doi: 10.4049/jimmunol.1001039 [DOI] [PubMed] [Google Scholar]
  • 109.Imamichi T, Yang J, Huang DW, et al. IL-27, a novel anti-HIV cytokine, activates multiple interferon-inducible genes in macrophages. Aids. 2008;22:39–45. doi: 10.1097/QAD.0b013e3282f3356c [DOI] [PubMed] [Google Scholar]
  • 110.Ilmarinen-Salo P, Moilanen E, Kankaanranta H. Nitric oxide induces apoptosis in GM-CSF-treated eosinophils via caspase-6-dependent lamin and DNA fragmentation. Pulm Pharmacol Ther. 2010;23:365–371. doi: 10.1016/j.pupt.2010.04.001 [DOI] [PubMed] [Google Scholar]
  • 111.Varin A, Mukhopadhyay S, Herbein G, Gordon S. Alternative activation of macrophages by IL-4 impairs phagocytosis of pathogens but potentiates microbial-induced signalling and cytokine secretion. Blood. 2010;115:353–362. doi: 10.1182/blood-2009-08-236711 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Joshi AD, Oak SR, Hartigan AJ, et al. Interleukin-33 contributes to both M1 and M2 chemokine marker expression in human macrophages. BMC Immunol. 2010;11:52. doi: 10.1186/1471-2172-11-52 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Venosa A, Malaviya R, Choi H, Gow AJ, Laskin JD, Laskin DL. Characterization of distinct macrophage subpopulations during nitrogen mustard-induced lung injury and fibrosis. Am J Respir Cell Mol Biol. 2016;54:436–446. doi: 10.1165/rcmb.2015-0120OC [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Ballantyne LL, Sin YY, Al-Dirbashi OY, Li X, Hurlbut DJ, Funk CD. Liver-specific knockout of arginase-1 leads to a profound phenotype similar to inducible whole body arginase-1 deficiency. Mol Genet Metab Rep. 2016;9:54–60. doi: 10.1016/j.ymgmr.2016.10.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Pauleau AL, Rutschman R, Lang R, Pernis A, Watowich SS, Murray PJ. Enhancer-mediated control of macrophage-specific arginase I expression. J Immunol. 2004;172:7565–7573. doi: 10.4049/jimmunol.172.12.7565 [DOI] [PubMed] [Google Scholar]
  • 116.Escamilla-Gil JM, Fernandez-Nieto M, Acevedo N. Understanding the cellular sources of the fractional exhaled nitric oxide (FeNO) and its role as a biomarker of type 2 inflammation in asthma. Biomed Res Int. 2022;2022:5753524. doi: 10.1155/2022/5753524 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Rupani H, Kent BD. Using fractional exhaled nitric oxide measurement in clinical asthma management. Chest. 2022;161:906–917. doi: 10.1016/j.chest.2021.10.015 [DOI] [PubMed] [Google Scholar]
  • 118.Sen P, Khatri SB, Tejwani V. Measuring exhaled nitric oxide when diagnosing and managing asthma. Cleve Clin J Med. 2023;90:363–370. doi: 10.3949/ccjm.90a.22072 [DOI] [PubMed] [Google Scholar]
  • 119.North ML, Grasemann H, Khanna N, Inman MD, Gauvreau GM, Scott JA. Increased ornithine-derived polyamines cause airway hyperresponsiveness in a mouse model of asthma. Am J Respir Cell Mol Biol. 2013;48:694–702. doi: 10.1165/rcmb.2012-0323OC [DOI] [PubMed] [Google Scholar]
  • 120.Nilsson BO, Hellstrand P. Effects of polyamines on intracellular calcium and mechanical activity in smooth muscle of Guinea-pig taenia coli. Acta Physiol Scand. 1993;148:37–43. doi: 10.1111/j.1748-1716.1993.tb09529.x [DOI] [PubMed] [Google Scholar]
  • 121.Ginguay A, Cynober L, Curis E, Nicolis I. Ornithine aminotransferase, an important glutamate-metabolizing enzyme at the crossroads of multiple metabolic pathways. Biology. 2017;6:18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Chakir J, Haj-Salem I, Gras D, et al. Effects of bronchial thermoplasty on airway smooth muscle and collagen deposition in asthma. Ann Am Thorac Soc. 2015;12:1612–1618. doi: 10.1513/AnnalsATS.201504-208OC [DOI] [PubMed] [Google Scholar]
  • 123.Shuhui L, Mok YK, Wong WS. Role of mammalian chitinases in asthma. Int Arch Allergy Immunol. 2009;149:369–377. doi: 10.1159/000205583 [DOI] [PubMed] [Google Scholar]
  • 124.Heyndrickx I, Deswarte K, Verstraete K, et al. Ym1 protein crystals promote type 2 immunity. Elife. 2024;12. doi: 10.7554/eLife.90676 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Zhao J, Zhu H, Wong CH, Leung KY, Wong WS. Increased lungkine and chitinase levels in allergic airway inflammation: a proteomics approach. Proteomics. 2005;5:2799–2807. doi: 10.1002/pmic.200401169 [DOI] [PubMed] [Google Scholar]
  • 126.Webb DC, McKenzie AN, Foster PS. Expression of the Ym2 lectin-binding protein is dependent on interleukin (IL)-4 and IL-13 signal transduction: identification of a novel allergy-associated protein. J Biol Chem. 2001;276:41969–41976. doi: 10.1074/jbc.M106223200 [DOI] [PubMed] [Google Scholar]
  • 127.Haeger SM, Yang Y, Schmidt EP. Heparan sulfate in the developing, healthy, and injured lung. Am J Respir Cell Mol Biol. 2016;55:5–11. doi: 10.1165/rcmb.2016-0043TR [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Dewald JH, Colomb F, Bobowski-Gerard M, Groux-Degroote S, Delannoy P. Role of cytokine-induced glycosylation changes in regulating cell interactions and cell signaling in inflammatory diseases and cancer. Cells. 2016;5. doi: 10.3390/cells5040043 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Zhao J, Lv Z, Wang F, et al. Ym1, an eosinophilic chemotactic factor, participates in the brain inflammation induced by Angiostrongylus cantonensis in mice. Parasitol Res. 2013;112:2689–2695. doi: 10.1007/s00436-013-3436-x [DOI] [PubMed] [Google Scholar]
  • 130.Dasgupta P, Chapoval SP, Smith EP, Keegan AD. Transfer of in vivo primed transgenic T cells supports allergic lung inflammation and FIZZ1 and Ym1 production in an IL-4Rα and STAT6 dependent manner. BMC Immunol. 2011;12:60. doi: 10.1186/1471-2172-12-60 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Raes G, De Baetselier P, Noël W, Beschin A, Brombacher F, Hassanzadeh Gh G. Differential expression of FIZZ1 and Ym1 in alternatively versus classically activated macrophages. J Leukoc Biol. 2002;71:597–602. doi: 10.1189/jlb.71.4.597 [DOI] [PubMed] [Google Scholar]
  • 132.Zhang L, Wang M, Kang X, et al. Oxidative stress and asthma: proteome analysis of chitinase-like proteins and FIZZ1 in lung tissue and bronchoalveolar lavage fluid. J Proteome Res. 2009;8:1631–1638. doi: 10.1021/pr800685h [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Doherty TA, Khorram N, Sugimoto K, et al. Alternaria induces STAT6-dependent acute airway eosinophilia and epithelial FIZZ1 expression that promotes airway fibrosis and epithelial thickness. J Immunol. 2012;188:2622–2629. doi: 10.4049/jimmunol.1101632 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Dong L, Wang SJ, Camoretti-Mercado B, Li HJ, Chen M, Bi WX. FIZZ1 plays a crucial role in early stage airway remodeling of OVA-induced asthma. J Asthma. 2008;45:648–653. doi: 10.1080/02770900802126941 [DOI] [PubMed] [Google Scholar]
  • 135.Liu T, Yu H, Ullenbruch M, et al. The in vivo fibrotic role of FIZZ1 in pulmonary fibrosis. PLoS One. 2014;9:e88362. doi: 10.1371/journal.pone.0088362 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Draijer C, Boorsma CE, Robbe P, et al. Human asthma is characterized by more IRF5+ M1 and CD206+ M2 macrophages and less IL-10+ M2-like macrophages around airways compared with healthy airways. J Allergy Clin Immunol. 2017;140:280–283.e283. doi: 10.1016/j.jaci.2016.11.020 [DOI] [PubMed] [Google Scholar]
  • 137.Lv R, Bao Q, Li Y. Regulation of M1‑type and M2‑type macrophage polarization in RAW264.7 cells by Galectin‑9. Mol Med Rep. 2017;16:9111–9119. doi: 10.3892/mmr.2017.7719 [DOI] [PubMed] [Google Scholar]
  • 138.Zhou Y, Do DC, Ishmael FT, et al. Mannose receptor modulates macrophage polarization and allergic inflammation through miR-511-3p. J Allergy Clin Immunol. 2018;141:350–364.e358. doi: 10.1016/j.jaci.2017.04.049 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Fricker M, Gibson PG. Macrophage dysfunction in the pathogenesis and treatment of asthma. Eur Respir J. 2017;50:1700196. doi: 10.1183/13993003.00196-2017 [DOI] [PubMed] [Google Scholar]
  • 140.Yang M, Kumar RK, Hansbro PM, Foster PS. Emerging roles of pulmonary macrophages in driving the development of severe asthma. J Leukoc Biol. 2012;91:557–569. doi: 10.1189/jlb.0711357 [DOI] [PubMed] [Google Scholar]
  • 141.Fitzpatrick AM, Teague WG, Burwell L, Brown MS, Brown LA. Glutathione oxidation is associated with airway macrophage functional impairment in children with severe asthma. Pediatr Res. 2011;69:154–159. doi: 10.1203/PDR.0b013e3182026370 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Liu MW, Liu R, Wu HY, et al. Atorvastatin has a protective effect in a mouse model of bronchial asthma through regulating tissue transglutaminase and triggering receptor expressed on myeloid cells-1 expression. Exp Ther Med. 2017;14:917–930. doi: 10.3892/etm.2017.4576 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Tsai ML, Tsai YG, Lin YC, et al. IL-25 induced ROS-mediated M2 macrophage polarization via AMPK-associated mitophagy. Int J Mol Sci. 2021;23:3. doi: 10.3390/ijms23010003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Liang Y, Shen S, Ye X, Zhang W, Lin X. Celastrol alleviates airway hyperresponsiveness and inflammation in obese asthma through mediation of alveolar macrophage polarization. Eur J Pharmacol. 2024;972:176560. doi: 10.1016/j.ejphar.2024.176560 [DOI] [PubMed] [Google Scholar]
  • 145.Tian C, Liu Q, Zhang X, Li Z. Blocking group 2 innate lymphoid cell activation and macrophage M2 polarization: potential therapeutic mechanisms in ovalbumin-induced allergic asthma by calycosin. BMC Pharmacol Toxicol. 2024;25:30. doi: 10.1186/s40360-024-00751-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Wang Y, Wan R, Peng W, Zhao X, Bai W, Hu C. Quercetin alleviates ferroptosis accompanied by reducing M1 macrophage polarization during neutrophilic airway inflammation. Eur J Pharmacol. 2023;938:175407. doi: 10.1016/j.ejphar.2022.175407 [DOI] [PubMed] [Google Scholar]
  • 147.Lin X, Zhang Y, Zhou X, Lai C, Dong Y, Zhang W. Inhibition of soluble epoxide hydrolase relieves adipose inflammation via modulating M1/M2 macrophage polarization to alleviate airway inflammation and hyperresponsiveness in obese asthma. Biochem Pharmacol. 2024;219:115948. doi: 10.1016/j.bcp.2023.115948 [DOI] [PubMed] [Google Scholar]
  • 148.He C, Chen P, Ning L, et al. Inhibition of mitochondrial succinate dehydrogenase with dimethyl malonate promotes M2 macrophage polarization by enhancing STAT6 activation. Inflammation. 2025. doi: 10.1007/s10753-024-02207-y [DOI] [PubMed] [Google Scholar]
  • 149.Fitzpatrick AM, Holguin F, Teague WG, Brown LA. Alveolar macrophage phagocytosis is impaired in children with poorly controlled asthma. J Allergy Clin Immunol. 2008;121:1372–1378, 1378.e1371–1373. doi: 10.1016/j.jaci.2008.03.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Song XM, Zhao MN, Li GZ, Li N, Wang T, Zhou H. Atorvastatin ameliorated myocardial fibrosis in db/db mice by inhibiting oxidative stress and modulating macrophage polarization. World J Diabetes. 2023;14:1849–1861. doi: 10.4239/wjd.v14.i12.1849 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Cattani-Cavalieri I, da Maia Valença H, Moraes JA, et al. Dimethyl fumarate attenuates lung inflammation and oxidative stress induced by chronic exposure to diesel exhaust particles in mice. Int J Mol Sci. 2020;21:9658. doi: 10.3390/ijms21249658 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Luo P, Liu D, Zhang Q, et al. Celastrol induces ferroptosis in activated HSCs to ameliorate hepatic fibrosis via targeting peroxiredoxins and HO-1. Acta Pharm Sin B. 2022;12:2300–2314. doi: 10.1016/j.apsb.2021.12.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Xia Y, Sun Y, Cao Y, et al. Calycosin alleviates sepsis-induced acute lung injury via the inhibition of mitochondrial ROS-mediated inflammasome activation. Front Pharmacol. 2021;12:690549. doi: 10.3389/fphar.2021.690549 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Sul OJ, Ra SW. Quercetin prevents LPS-induced oxidative stress and inflammation by modulating NOX2/ROS/NF-kB in lung epithelial cells. Molecules. 2021;26:6949. doi: 10.3390/molecules26226949 [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Journal of Asthma and Allergy are provided here courtesy of Dove Press

RESOURCES