Next Article in Journal
Complement Modulation Mitigates Inflammation-Mediated Preterm Birth and Fetal Neural Inflammation
Previous Article in Journal
The Integrated Function of the Lateral Hypothalamus in Energy Homeostasis
Previous Article in Special Issue
Role of Oxidative Stress and Inflammation in Postoperative Complications and Quality of Life After Laryngeal Cancer Surgery
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

p47phox: A Central Regulator of NADPH Oxidase Function and a Promising Therapeutic Target in Redox-Related Diseases

1
Department of Medicine, University of Arizona Cancer Center, Tucson, AZ 85724, USA
2
Department of Pharmacology and Toxicology, University of Arizona College of Pharmacy, Tucson, AZ 85721, USA
3
Division of Hematology, University of Southern California Keck School of Medicine, Los Angeles, CA 90033, USA
*
Author to whom correspondence should be addressed.
Cells 2025, 14(14), 1043; https://doi.org/10.3390/cells14141043
Submission received: 8 May 2025 / Revised: 2 July 2025 / Accepted: 7 July 2025 / Published: 8 July 2025

Abstract

The NADPH oxidase 2 (NOX2) complex is a critical regulator of immune homeostasis. It is utilized by phagocytic leukocytes including neutrophils, monocytes, and macrophages to generate reactive oxygen species (ROS) that drive microbe clearance and modulate inflammatory responses. Within NOX2, the essential scaffold protein p47phox plays a pivotal role in orchestrating enzyme activation and facilitating the assembly and membrane translocation of cytosolic components of the complex. Tight regulation of p47phox activity is crucial, and its disruption is linked to a number of pathological conditions. Conversely, its hyperactivity contributes to oxidative stress, tissue damage, the progression of cardiovascular diseases, neurodegenerative disorders, inflammatory conditions, metabolic syndromes, and cancer. In this review, we detail the structural and functional roles of p47phox, mechanisms of its regulation, and its multifaceted contributions to disease pathogenesis. We explore the latest advances in p47phox-targeted therapeutic strategies, discuss current challenges in the field, highlight p47phox’s potential as a transformative target in redox biology and propose future directions to unlock its clinical utility.

1. Introduction

Reactive oxygen species (ROS) are critical for the maintenance of immune homeostasis and microbial killing within the phagosome. Phagocytic leukocytes (neutrophils, eosinophils, monocytes, and macrophages) utilize pattern recognition receptors (PRRs) and toll-like receptors (TLRs) to detect pathogens. This primes the NADPH oxidase (NOX) enzyme complex, enhancing its capacity to mount a defensive respiratory burst upon subsequent activation by stimuli such as opsonized microbes or bacterial peptides [1,2]. Superoxide anions (O2) generated by NOX activity are rapidly converted into hydrogen peroxide (H2O2), hydroxyl radicals (OH•), and hypochlorous acid (HOCl) via secondary enzymatic reactions [3,4]. This process is normally tightly controlled, but its dysregulation can lead to oxidative damage to host tissues, chronic inflammation, cellular senescence, and a wide variety of diseases [5,6].
Phagocytes preferentially utilize the NADPH oxidase complex to generate reactive oxygen species (ROS) for host defense. This multicomponent enzyme system—often referred to as NOX2—includes the membrane-bound catalytic core (gp91phox, also known as NOX2, and p22phox), cytosolic regulatory subunits (p47phox, p67phox, p40phox) and regulatory GTPases such as Rac1/2 [7,8,9]. The scaffold protein p47phox (neutrophil cytosolic factor 1, NCF1) plays a central role in NOX2 activation through its conformational dynamics. When at rest, it adopts an autoinhibited conformation that prevents unwanted basal ROS production [10]. Invading pathogens and inflammatory signals trigger the serine phosphorylation of p47phox (Ser303–Ser379), which induces a conformational change that exposes its phox homology (PX) and Src homology 3 (SH3) domains. This facilitates membrane translocation and assembly of the NOX2 complex [11].
Given p47phox’s role as a key determinant of NOX2 activity, it is not surprising that its dysfunction has been linked to immunodeficiency and hyperinflammatory states [12]. For example, mutations or deletions in NCF1 cause a rare genetic disorder called chronic granulomatous disease (CGD) that is associated with defective ROS production, recurrent infections, and granuloma formation [13,14]. On the flip side, excessive p47phox-driven oxidative stress contributes to atherosclerosis, Parkinson’s disease, Alzheimer’s, rheumatoid arthritis (RA), diabetes, cancer, etc. [6,9,15]. p47phox’s important role in so many diseases creates opportunities for innovative precision therapeutic strategies to inhibit or activate its function where appropriate.
This review provides an in-depth analysis of p47phox’s structural architecture, its mechanisms of action and contributions to various disease states. We highlight recent advances in p47phox-targeted therapies, address current challenges in the field and propose future directions for research and clinical translation. Drawing on the latest literature, we position p47phox as a key player in redox biology and an emerging target for precision medicine in oxidative stress-related disorders.

2. Structural Features of p47phox

p47phox is a 390-amino-acid protein encoded by the NCF1 gene located on chromosome 7q11.23. It is highly abundant in neutrophils (100–150 ng per million cells) to ensure rapid NOX2 activation during immune challenges [16]. The structure of p47phox is comprised of multiple functional domains, with each contributing to its regulatory and interactive roles within the NOX2 complex (Figure 1). The N-terminal PX domain (residues 4–121) facilitates membrane association during NOX2 activation through specific binding to phosphoinositides, particularly phosphatidylinositol 3,4-bisphosphate (PtdIns(3,4)P2). Unlike the PX domain of p40phox, which binds phosphatidylinositol 3-phosphate (PtdIns(3)P), the PX domain of p47phox shows minimal affinity for PtdIns(3)P, but may also interact with phosphatidylserine (PS) or phosphatidic acid (PA) [17,18,19]. The PX domain is followed by two tandem Src homology 3 (SH3) domains that enable protein–protein interactions that are essential for NOX2 function. The SH3A (residues 159–214) and SH3B (229–284) domains stabilize the membrane-bound NOX2 core by binding to the proline-rich region (PRR) of p22phox [20,21]. Mutations within this region disrupt p47phox–p22phox interactions and severely impair superoxide production. Key examples include the Δ219–222 deletion (reduces superoxide generation by up to 60%) and the W193R substitution (completely inhibits superoxide production) [22,23]. These findings highlight the structural and functional importance of the p47phox SH3 domains as regulators of NADPH oxidase activity. The C-terminal autoinhibitory region (AIR, residues 292–340) maintains p47phox in a closed, inactive conformation by masking its interaction domains to prevent premature NOX2 activation [24]. Finally, the C-terminal proline-rich region (PRR, residues 363–368) interacts with the SH3 domain of p67phox to facilitate the pre-assembly of the cytosolic complex (p47phox-p67phox-p40phox) before membrane recruitment [25,26].
Structural insights into p47phox have primarily been gained through X-ray crystallography and nuclear magnetic resonance (NMR) studies, which have elucidated the conformation of its key domains in both autoinhibited and active states. The PX domain structure was resolved by X-ray crystallography and revealed a conserved phosphoinositide-binding pocket critical for membrane targeting [19]. The tandem SH3 domains have been characterized by crystallography, showing a compact fold that enables simultaneous interaction with the p22phox PRR and regulatory flexibility [24]. Small-angle X-ray scattering (SAXS) and hydrogen-deuterium exchange mass spectrometry (HDX-MS) have further contributed to understanding the conformational dynamics of full-length p47phox during phosphorylation-induced activation [27]. While a high-resolution structure of the full-length protein remains unavailable, integrative modeling and molecular dynamics simulations have provided evidence for domain rearrangements and allosteric regulation, particularly involving the autoinhibitory AIR domain and its release upon serine phosphorylation [28]. These studies collectively underscore the modular and dynamic nature of p47phox, which enables precise spatiotemporal regulation of NOX2 activity in immune responses. Moreover, these efforts identified transient allosteric sites that could potentially be targeted for therapeutic intervention [27,28,29].

3. Molecular Mechanisms of p47phox in NOX2 Activation

3.1. Phosphorylation and Conformational Change

p47phox adopts a compact, autoinhibited conformation in its resting state [10]. NOX2 activation occurs at phagosomal or plasma membranes where the cytochrome b558 complex (gp91phox and p22phox) is located. Activation is highly regulated and involves the transfer of electrons from NADPH to molecular oxygen to generate superoxide [1,5]. In resting neutrophils, the cytosolic components (p47phox, p67phox, p40phox, and Rac2) remain sequestered in the cytoplasm to maintain NOX2 in an inactive state. This prevents basal ROS production and oxidative damage to host tissues [7]. Priming of phagocytes is initiated by pathogen-associated molecular patterns (PAMPs), damage-associated molecular patterns (DAMPs), or inflammatory cytokines. This prepares the NADPH oxidase complex for rapid activation upon subsequent stimulation by agents such as the bacterial peptide formyl-methionyl-leucyl-phenylalanine (fMLF).
The functional activity of p47phox is tightly regulated by a complex network of phosphorylation events, PTMs and protein–protein interactions (Table 1). This process is essential for redox homeostasis and ensures precise control of NOX2 activation and ROS production. Activation begins with phosphorylation of its C-terminal serine-rich region (Ser303–Ser379), which acts as a molecular switch for NOX2 assembly [30]. This can be mediated by PKC isoforms (α, βII, δ, ζ), ERK1/2, Akt, p38 MAPK, and others [31]. Phosphorylation of key residues such as Ser379 within the conserved “RRXSXR” motif induces a conformational shift that disrupts the AIR to expose the PX and SH3 domains. This enables p47phox to bind PtdIns(3,4)P2 at the membrane and interact with p22phox via its SH3 domains [10,20]. The resulting conformational change transforms p47phox from a compact, functionally inactive state into an extended, functionally active form that facilitates the recruitment of p67phox and p40phox to form the active NOX2 complex (Figure 2) [32]. Earlier studies showed that treatment with phorbol 12-myristate 13-acetate (PMA) rapidly leads to Ser379 phosphorylation-driven NOX2 activation [32]. Additional in vitro investigations demonstrated that PKC-mediated phosphorylation of recombinant p47phox in the presence of p67phox and Rac2 was sufficient to trigger an oxidative burst. This study validated its role as a master regulator of the complex [30]. While Ser379 is the most well-characterized functional activation-linked p47phox phosphorylation site, recent phosphoproteomic analyses have identified additional potential sites for PKC-driven functional activation, including Ser345. Initial evidence suggests that phosphorylation of these less intensively studied sites may enhance p47phox membrane affinity under specific inflammatory conditions [33]. Additional studies are required to better define these dynamics.
The current working model of kinase-driven NOX2 regulation involves PKC driving the rapid, acute phase of ROS production with AKT and MAPKs sustaining and amplifying the response over time. This coordinated interplay ensures a balanced and adaptable oxidative response. Evidence suggests that AKT-driven phosphorylation of p47phox on Ser304 and Ser328 stabilizes its active conformation to promote sustained ROS generation [34]. p38 MAPK and ERK1/2 can also modulate NOX2 activity via Ser345 phosphorylation. This enhances p47phox responsiveness to secondary stimuli, such as fMLF. Pro-inflammatory cytokines like TNF-α act as priming agents, amplifying NOX2-dependent ROS production during subsequent microbial encounters. This mechanism is heavily utilized by neutrophils for antimicrobial defense [33].
The kinase-related regulatory landscape of NOX2 activity is further complemented by protein kinase A (PKA) and casein kinase 2 (CK2). PKA has been specifically implicated in phosphorylating p47phox at distinct sites (e.g., Ser320) in response to cyclic AMP signaling [35,36]. CK2 may also modulate p47phox functional activity under specific stress conditions [37]. The complexity of kinase-mediated regulation of p47phox underscores its role as a molecular rheostat in NOX2 activation. Indeed, excessive kinase activity or impaired dephosphorylation have both been linked to chronic inflammatory diseases and oxidative stress-related disorders. The improved understanding of how this kinase network works has highlighted clear opportunities for precision modulation of ROS production.

3.2. Role of Cytosolic Regulators

Once phosphorylated, p47phox serves as a scaffold that recruits p67phox and p40phox to membrane-bound cytochrome b558. The C-terminal PRR of p47phox binds the SH3 domain of p67phox. The complex is then stabilized through the binding of p40phox to the PB1 domain of p67phox [18,24]. NOX2 stability and catalytic efficiency are further enhanced when GTP-bound active Rac2 binds to gp91phox and p67phox. This ensures robust electron transfer and superoxide production [34]. Recent studies using fluorescence resonance energy transfer (FRET) have mapped these interactions in real time. Molecular dynamics simulations and AlphaFold-predicted structures have provided further insights into the assembly and activation of the NOX2 complex [28,38]. These findings support a model in which cytosolic subunits (p47phox, p67phox, and p40phox) translocate as a pre-assembled complex, while Rac2 is recruited independently and plays a key role in stabilizing and orienting the NOX2 complex at the membrane.

3.3. Efficiency of NOX2 Activation

Although only a fraction of cytosolic p47phox actually translocates to the membrane upon stimulation, this is sufficient to trigger substantial superoxide generation [8]. This efficiency is attributed to the synergistic dynamics between p47phox, p67phox and Rac2 in NOX2 assembly and stabilization. Mutagenesis studies have confirmed the essential role of p47phox in the NOX2 activation process. For example, a serine to alanine mutation at position 379 completely impairs NOX2 activity. Dual Ser303/Ser304 mutations dramatically diminish superoxide production [30]. These collective findings underscore the importance of these specific phosphorylation sites as regulators of the complex. Additionally, computational modeling and inhibitor studies (e.g., C6 and C14) have identified key molecular interfaces, including gp91phox regions (residues 86–93, 450–457), that indirectly influence NOX2 regulation via p22phox. These interfaces could represent potential targets for therapeutic intervention [39,40].

3.4. Role of Pin1

The interaction of prolyl isomerase Pin1 with phospho-Ser345 of p47phox induces a conformational change that exposes additional serine residues for PKC phosphorylation to amplify NOX2 activity [19]. This is another example of the regulatory feedback loop of phosphorylation-dependent structural changes that titrate p47phox function. Recent studies suggest that Pin1 promotes p47phox activation by facilitating conformational changes necessary for NOX2 complex assembly. This interaction plays an important regulatory role and may represent a potential therapeutic target for modulating ROS production [41]. Additional layers of regulation are achieved through the binding of 14-3-3 proteins to phosphorylated p47phox. This binding stabilizes its active conformation, while its dephosphorylation by the PP2A phosphatase inhibits NOX2 activity [42].

3.5. Other Post-Translational Modifications

In addition to phosphorylation, p47phox is regulated by several other post-translational modifications (PTMs) that influence its stability, interactions, and activity within the NOX2 complex. For example, p47phox levels can be titrated during prolonged immune responses to modulate ROS production through lysine-targeted ubiquitination to degrade a portion via the ubiquitin–proteasome system (UPS) [43]. Conversely, lysine acetylation mediated by enzymes like p300 offers another mechanism to control p47phox stability and promote its association with other NOX2 subunits under oxidative stress conditions [44]. ROS-triggered S-glutathionylation of cysteine residues may protect p47phox from irreversible oxidative damage or fine-tune its activation in response to redox stress [45]. Additionally, S-nitrosylation has been proposed as a mechanism by which nitric oxide signaling could inhibit NOX2 activity through p47phox modification [46]. Finally, O-GlcNAcylation has emerged as a potential regulatory mechanism linking metabolic cues to p47phox function [47]. Collectively, these PTMs underscore the multifaceted regulation of p47phox and present opportunities for therapeutic intervention in redox-driven diseases.

4. p47phox in Pathological Conditions

4.1. Chronic Granulomatous Disease (CGD)

Chronic granulomatous disease (CGD) is a rare primary immunodeficiency. It affects approximately 1 in 200,000 individuals worldwide and is characterized by impaired ROS production that results from NOX2 defects [13]. NCF1 mutations account for approximately 25% of all CGD cases [48]. The most common NCF1 mutation in patients with CGD is a deletion of GTGT in exon 2. This deletion results in a frameshift that prematurely terminates the protein and impairs superoxide generation [49]. Consequently, phagocytes are left unable to kill pathogens such as Staphylococcus aureus, Aspergillus, Burkholderia cepacia, and Nocardia. Affected patients experience recurrent, life-threatening infections [50]. The severe immunodeficiency caused by NCF1 defects translates into most CGD patients being diagnosed at an early age. They typically begin experiencing abscesses, pneumonia, and fungal infections along with chronic inflammation-induced granuloma formation by age 10 [51]. The increased susceptibility of CGD patients to infections is linked to the redox-related downstream consequences of p47phox deficiency. The impaired ability to generate superoxide means that these patients also produce very low levels of hydrogen peroxide and hypochlorous acid, both of which are critical for microbial killing [52]. Beyond their primary neutrophil defects, CGD patients also experience additional immune dysfunction due to impaired macrophage antigen presentation [53]. The complexities of the immunobiology of CGD have been defined largely through studies with NCF1/ mice, which faithfully replicate the primary features of human CGD. Like humans with CGD, NCF1/ mice exhibit granuloma formation, increased IL-1β and TNF-α production, and heightened susceptibility to Aspergillus infections [54]. A recent study linked p47phox deficiency in CGD to increased microbiota-related susceptibility to colitis. The investigators demonstrated that intestinal epithelial cells from NCF1/ mice produced increased levels of ROS as compared with gp91phox/− mice. Their findings suggest that the susceptibility of CGD patients to colitis development is driven in a genotype-specific manner [55]. Not only does the ability to model the disease in mice offer an opportunity to mechanistically investigate the complex physiology of CGD, it also provides a platform to preclinically assess novel therapeutics to better manage the disease (Table 2).
Current treatment strategies for CGD focus on improving immune defense through prophylactic antibiotics, antifungals, and interferon-γ, but do not correct the underlying deficiency in ROS production that drives disease pathology [54]. Gene therapy using lentiviral vectors to deliver functional NCF1 has shown promise, achieving partial restoration of ROS production in patient-derived neutrophils and a reduction in infection rates in preclinical models [56]. Clinical investigation of the safety and efficacy of this approach has only recently been initiated [57].

4.2. Cardiovascular Diseases

p47phox-driven oxidative stress has emerged as a central player in cardiovascular diseases including atherosclerosis, hypertension, and endothelial dysfunction [58]. The production of ROS in the vascular wall sets off a cascade of damage. The oxidization of low-density lipoprotein (LDL) cripples nitric oxide (NO) signaling and sparks the inflammation that ultimately leads to plaque buildup [59,60]. The accumulation of LDL particles in arterial walls transforms macrophages into foam cells that pile up lipids and form the fatty streaks of early atherosclerosis [61]. This process ultimately culminates in the hardening and narrowing of arteries.
Studies in mice have painted a striking picture of the true impact of p47phox on cardiovascular pathogenesis. The infusion of angiotensin II into mice lacking p47phox expression results in a significant drop in vascular superoxide levels compared to their wild-type counterparts. Their systolic blood pressure returned to the near normal range and their atherosclerotic plaques also shrank [62,63]. These findings suggest that p47phox acts as an architect of vascular damage in hypertension and atherosclerosis. The increased basal ROS levels in endothelial cells that are triggered by p47phox accelerate plaque formation [64]. This mechanistic relationship offers an opportunity for therapeutic intervention. Antioxidants like Ginsenoside Rb1 can reduce superoxide production and improve vascular function [65]. While not curative, this type of approach could be helpful in controlling disease progression. Additionally, a recent study linking the lysine methyltransferase SET7 to NOX-driven atherosclerosis suggests that epigenetic targeted therapies might also aid in managing the disease [66].
NO is critically important for keeping blood vessels structurally flexible. This prevents monocytes from sticking and triggering inflammation. Unfortunately, p47phox hyperactivity antagonizes the benefit of NO in this context. When p47phox-driven ROS oxidizes NO or its precursors, it reduces NO availability. This leads to stiffened vessels and the recruitment of inflammatory cells [67]. Deletion of p47phox (NCF1) in the ApoE knockout mouse model of atherosclerosis reduces aortic ROS levels and shrinks plaque size [68]. These findings demonstrate the pathological significance of unchecked p47phox activity on cardiovascular homeostasis. Indeed, a study of patients with advanced atherosclerotic lesions revealed that they often have higher levels of NCF1 mRNA in their plaques [69]. There is also evidence that NCF1 gene variants may contribute to hypertension by influencing ROS production in vascular cells. However, their specific impact on stroke risk remains under investigation [70].
The role of p47phox in cardiovascular disease is ultimately one of balance. While it is essential for fighting pathogens, it can be incredibly destructive when functionally hyperactive in blood vessels. As we continue to define its role in cardiovascular pathogenesis, we will be able to harness that knowledge to design precision therapies. Inhibiting it could antagonize the oxidative stress that drives cardiovascular damage, but care needs to be taken to ensure that doing so does not compromise immune function. Modulating upstream pathways like angiotensin II signaling or glucose metabolism that amplify p47phox may offer an alternative approach [71]. There is also growing interest in the potential use of novel antioxidants or compounds that stabilize NO signaling to complement p47phox modulation [72]. Further investigation is required to ultimately determine which strategy is optimal for individual patients. This will likely depend largely on their overall health status as comorbidities could significantly change the landscape.

4.3. Neurodegenerative Diseases

p47phox has emerged as a driver of oxidative stress, neuroinflammation and neuronal loss in conditions like Parkinson’s disease (PD) and Alzheimer’s disease (AD) [73]. By orchestrating excessive ROS production in activated microglia, p47phox initiates a cascade that damages the delicate architecture of the brain. p47phox is also linked to the destruction of dopaminergic neurons in the substantia nigra, the region of the brain that is critical for motor control in PD. p47phox-mediated ROS also promote the aggregation of α-synuclein, which is a hallmark feature of PD. This oxidative onslaught also disrupts mitochondrial homeostasis by triggering membrane depolarization, cytochrome c release, and neuronal cell apoptosis [74,75]. Studies using MPTP (1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine), a toxin that mimics PD in mice, demonstrated that mice lacking the NCF1 gene retain significantly more dopaminergic neurons than their wild-type counterparts. They also exhibit superior motor coordination and lower levels of oxidative damage markers such as 4-hydroxynonenal (4-HNE) [76]. These studies suggest that inhibiting p47phox activity may be sufficient to antagonize PD progression. p47phox also contributes to the pathogenesis of AD through other mechanisms. ROS generated by p47phox-activated microglia trigger the hyperphosphorylation of the tau protein and amyloid beta (Aβ) plaque buildup [77]. These alterations cause synaptic dysfunction and erode cognitive abilities. In the APP/PS1 mouse model of AD, knocking out p47phox is sufficient to limit tau pathology [78]. These findings suggest that p47phox dysfunction plays a pivotal role in AD pathogenesis [79].
The therapeutic potential of targeting p47phox for the treatment of neurodegenerative disorders is scientifically rational. Small molecule NOX2 inhibitors like Ebselen have shown promise in preclinical models of PD and AD. Their ability to reduce oxidative stress leads to an increase in in vitro neuronal cell survival and reduced inflammation and neuronal loss in mouse models of disease [80]. However, the safety and efficacy of long-term p47phox inhibition remains uncertain. Since NOX2 helps to clear debris and fight infections in normal microglial cells, it is possible that chronic suppression could disrupt brain homeostasis or increase vulnerability to pathogens [81]. In view of this, alternative strategies to counter the pathogenic effects of p47phox are being explored. Modulation of upstream signals that amplify p47phox activity like angiotensin II and inflammatory cytokines are being investigated in neurodegenerative contexts [82]. Others are evaluating the potential benefit of antioxidants tailored to neutralize p47phox-derived ROS without broadly impairing NOX2 [83]. A deeper understanding of the mechanisms by which p47phox interacts with α-synuclein, tau, and Aβ may facilitate a balanced strategy that does not compromise the brain’s defenses.

4.4. Inflammatory Disorders

p47phox plays a pivotal yet dichotomous role in the pathogenesis of inflammatory disorders. In rheumatoid arthritis (RA), p47phox drives synovial inflammation and joint destruction. Excessive p47phox-mediated NOX2 activation generates a surge of ROS within the synovium, which causes oxidative damage to cartilage and bone [84,85]. The relationship between p47phox and disease activity has been validated in synovial biopsies from RA patients, which demonstrated elevated p47phox expression and significant superoxide generation [86]. Additional studies have shown that the peptidyl-prolyl isomerase Pin1 enhances p47phox phosphorylation to amplify NOX2 activity in RA [87]. This hyperactive state sustains ROS production while also activating the NF-κB signaling pathway to drive the expression of pro-inflammatory cytokines like interleukin-6 (IL-6) and TNF-α (Table 3) [88]. Preclinical studies in collagen-induced arthritis models established proof of concept for therapeutic Pin1 inhibition. Synovial ROS levels were reduced, cartilage erosion was limited and NF-κB-driven inflammation was suppressed in response to Pin1 inhibition [89]. These findings suggest that targeted inhibition of p47phox or its upstream regulators could offer novel avenues for disease management.
Conversely, p47phox plays a protective role in inflammatory bowel disease (IBD) by maintaining gut homeostasis. In the intestinal mucosa, ROS contribute to microbial killing and the reinforcement of the integrity of the epithelial barrier. Both are critical defenses against dysbiosis and inflammation [90]. Studies in NCF1/ mice revealed that p47phox deficiency resulted in increased susceptibility to colitis in a manner that was directly influenced by the microbiome [91]. This provided further evidence that p47phox activity is required for gut immune homeostasis [92]. Reduced NCF1 expression is also associated with more frequent flares in patients with Crohn’s disease. This is likely due to impaired microbial control. Similarly, elevated NCF1 expression is correlated with greater disease severity in patients with ulcerative colitis. This has been attributed to excessive ROS-driven inflammation [93]. Collectively, these studies illuminate the paradoxical role of p47phox in gut immunity, where both under- and overactivity can precipitate pathology. Beyond RA and IBD, emerging evidence implicates p47phox in other inflammatory conditions, such as systemic lupus erythematosus (SLE) and psoriasis. In SLE, NCF1 polymorphisms are linked to increased disease susceptibility due to p47phox-driven ROS-mediated autoantigen generation and immune complex-mediated tissue damage [94]. Similarly, in psoriasis, p47phox activation in keratinocytes amplifies ROS-dependent inflammatory signaling. This exacerbates epidermal hyperplasia [95]. These examples further illustrate the context-specific nature of p47phox’s contributions to inflammation.
Therapeutically, modulating p47phox activity in inflammatory diseases presents both opportunities and challenges. In RA, inhibitors targeting NOX2 or its upstream activators like Pin1 or PKC may offer an opportunity to reduce synovial oxidative stress without broadly suppressing immunity. Small molecule NOX2 inhibitors have shown efficacy in reducing joint inflammation in preclinical models of the disease [96]. However, their specificity and long-term safety require further validation. In contrast, approaches to enhance p47phox function in IBD like Crohn’s disease could potentially restore microbial homeostasis. To be successful, such strategies would need to avoid triggering excessive ROS production as that could exacerbate inflammation. Precision medicine approaches guided by genetic profiling of NCF1 variants or tissue-specific ROS profiling could help tailor interventions to individual disease contexts. The current evidence in the field indicates that p47phox acts as a molecular fulcrum that balances protective and pathological outcomes. As continued research further defines the molecular and contextual determinants of p47phox activity, precision therapeutics that fine-tune ROS production in a disease-specific manner may become a reality.

4.5. Metabolic Disorders

p47phox-driven ROS play a pivotal role in metabolic diseases such as diabetes mellitus by driving the underlying oxidative stress and tissue damage that contribute to pathogenesis in multiple organ sites. In type 1 and type 2 diabetes, hyperglycemia activates p47phox to amplify ROS production via multiple signaling pathways including PKC, MAPK, and the renin–angiotensin system. This impairs insulin signaling, glucose uptake, and mitochondrial function [97]. The resulting oxidative burden disrupts redox homeostasis and promotes lipid peroxidation, protein carbonylation, and DNA damage in pancreatic β-cells, adipocytes, hepatocytes, and vascular endothelial cells. Ultimately, this exacerbates insulin resistance and triggers β-cell apoptosis [98].
Diabetic retinopathy is a leading cause of blindness in diabetic patients. It has also been linked to p47phox-driven oxidative stress in retinal endothelial cells and pericytes. Hyperglycemia-induced p47phox activation has been shown to double vascular ROS levels and trigger endothelial dysfunction, pericyte loss, and neovascularization [99]. The related retinal pathology has been tied to the induction of vascular endothelial growth factor (VEGF) and matrix metalloproteinases (MMPs), which heighten vascular permeability and angiogenesis [100]. In mouse models of diabetic retinopathy, photoreceptor-driven oxidative stress via NOX contributes to ROS generation, retinal vascular leakage, and capillary degeneration. This suggests a potential role for p47phox in these pathologies and indicates that therapeutic p47phox inhibition may hold promise as a novel approach to preserve vision in diabetic patients [101].
p47phox hyperactivation is also a contributing factor in diabetic renal dysfunction. In glomerular mesangial cells and podocytes, increased ROS levels have been linked to extracellular matrix accumulation, fibrosis, and albuminuria [102]. p47phox-mediated ROS production has been shown to mediate these pathogenic effects through the activation of TGF-β and NF-κB signaling pathways. This promotes the pro-fibrotic and pro-inflammatory responses that underlie glomerulosclerosis [103]. In further support of this deleterious relationship, studies in NCF1/ diabetic mouse models demonstrated that p47phox deficiency results in a reduction in glomerular ROS, decreased albuminuria, and improved renal function [104]. p47phox-mediated ROS generation in peripheral nerves and Schwann cells is also believed to contribute to diabetic neuropathy by promoting axonal degeneration and sensory loss. Mechanistically, p47phox has been hypothesized to promote neuropathy in this context through ROS-mediated activation of poly(ADP-ribose) polymerase (PARP) and caspase-3, which stimulates neuronal apoptosis and mitochondrial dysfunction [99].
Although the scientific support for a role for p47phox in the pathogenesis of metabolic disorders is currently strongest in diabetes, there is growing evidence that it also contributes to the development or progression of other metabolic diseases. Non-alcoholic fatty liver disease (NAFLD) and obesity are two examples of this where ROS generation in hepatic and adipose tissue drive insulin resistance and steatosis [105,106,107]. Collectively, these findings indicate that tightly controlled p47phox activity is required to maintain metabolic homeostasis. The clear role for p47phox as a driver of multiple complications that are associated with metabolic dysfunction highlights its potential significance as a therapeutic target. However, potential challenges for the successful development of targeted agents include off-target effects in normal tissues and the possible need for tissue-specific delivery systems to prevent immune suppression. Ongoing research using advanced omics approaches and sophisticated preclinical models will empower the development of the best strategies for precision targeting in different disease scenarios.

4.6. Cancer

p47phox plays divergent roles in cancer. It can promote tumor progression via ROS-mediated oncogenic signaling while also facilitating antitumor immunity through immune cell activation. This duality reflects its context-dependent expression and activity in tumor cells, stromal cells, and immune infiltrates. These factors make it a complex yet promising therapeutic target [108]. p47phox-driven ROS in malignant cells sustain signaling through the PI3K/AKT and MAPK pathways to drive proliferation, survival, metastasis, and chemoresistance [109]. Conversely, in immune cells, p47phox-derived ROS enhance cytotoxic T-cell and macrophage responses against tumors. Studies in NCF1/ mice suggest that p47phox deficiency modulates immune responses in cancer, potentially affecting tumor progression through altered ROS signaling [110]. These opposing effects suggest that targeted p47phox modulation could potentially enhance immunotherapy while suppressing tumor-promoting ROS.
The current evidence suggests that p47phox activity has both recurrent and distinct consequences across different malignancies. In melanoma, elevated NCF1 expression correlates with increased basal ROS generation, enhanced BRAF/MEK/ERK signaling and elevated metastatic burden [111,112]. In estrogen receptor-positive (ER+) and triple-negative breast cancer cells, data indicate that p47phox amplifies ROS production to sustain HER2/Neu and EGFR signaling [113]. In mammary tumor models, NCF1/ may reduce tumor progression through altered NOX activity by affecting MMP-9 and VEGF [114]. p47phox also contributes to chemoresistance and studies suggest that it and other NOXs promote aggressive cancer pathophysiology through ROS-mediated signaling [114,115,116,117].
The Wnt/β-catenin signaling cascade is an important driver of epithelial-to-mesenchymal transition (EMT) and metastasis in colorectal cancer. Evidence suggests that p47phox-driven ROS regulate both of these processes [118]. Knockout of the NCF1 gene in the ApcMin mouse model results in fewer intestinal adenomas and reduced basal ROS levels [110]. p47phox has also been shown to activate NF-κB in tumor-associated macrophages, which promotes chronic inflammation and tumor progression [119]. However, its role in immune system dynamics is complex. p47phox has also been shown to boost ROS in neutrophils, macrophages, and CD8+ T cells, which could enhance antitumor immunity [120].
The tumor microenvironment (TME) also plays a role in modulating p47phox activity. Hypoxia and cytokine signaling have been shown to amplify its activity in stromal fibroblasts and endothelial cells. This has been linked to angiogenesis and metastasis [121,122,123,124]. Studies conducted in preclinical models of pancreatic cancer demonstrated that p47phox increases ROS generation in stromal cells and promotes desmoplasia and chemoresistance. These findings support its potential promise as a therapeutic target [125,126]. However, the ability of p47phox modulation to yield therapeutic benefit appears to be malignancy-specific and context dependent. The known immunomodulatory effects of p47phox redox-driven signaling indicate that there may be an opportunity for synergy between p47phox stimulators and immune checkpoint inhibitors [127]. Future studies focused on defining the intricate roles of p47phox in malignant pathogenesis will enable the development of optimized therapeutic strategies.

5. Therapeutic Potential of Targeting p47phox

5.1. p47phox Inhibitors

Inhibiting p47phox is a rational therapeutic strategy to reduce oxidative stress for diseases where excessive ROS generation drives pathology, including cardiovascular disease, RA, neurodegeneration, metabolic disorders, and cancer. To date, the development of p47phox inhibitors has largely focused on disrupting its protein–protein interactions (PPIs) within the NOX2 complex (Table 4, Figure 3). Targeting the p47phox–p22phox interface has been a major focus as this approach may reduce the probability of off-target effects on other NOX isoforms or cellular pathways [128]. Peptide-based inhibitors like NOX2ds-tat have emerged as highly specific tools that mimic the PRR of p22phox to block p47phox’s SH3 domain interactions [129,130]. In a study conducted in angiotensin II-induced hypertensive mice, NOX2ds-tat treatment reduced vascular ROS, lowered systolic blood pressure and significantly decreased the atherosclerotic plaque burden [131]. It also demonstrated therapeutic activity with respect to vascular compensation in a model of diet-induced obesity that mimicked a diabetic phenotype [132]. Similarly, NOX2ds-tat improved inflammation following traumatic brain injury, as well as pain-related behavior associated with spinal cord injuries, in preclinical studies [133,134]. Despite these encouraging results, there are challenges that may ultimately limit the successful clinical development of NOX2ds-tat and other peptide inhibitors including rapid renal clearance, limited intracellular penetration, susceptibility to proteolytic degradation, and lack of isoform specificity [129,130,135,136]. Further structure modification and innovative formulations could potentially improve its drug-like properties and make it a stronger candidate for clinical testing, but additional research is needed to determine this.
The current consensus is that small molecule inhibitors are likely to offer greater clinical feasibility over peptide-based targeting. CPP11G and CPP11H are two small molecule compounds that were designed to disrupt p47phox–p22phox binding. A preclinical study conducted with these agents in a vascular dysfunction model showed that they enhanced vascular function by lowering endothelial ROS, improving NO bioavailability, and restoring hind-limb blood flow [137]. As was learned from the preclinical evaluation of NOXA2ds-tat, there are also challenges that will likely limit the potential clinical development of CPP11G and CPP11H. These include poor solubility, potential selectivity issues, high dosing requirements and challenges inherent to PPI inhibitors. However, additional studies are required to determine if advancement into human clinical trials is warranted.
Celastrol and Ebselen are two additional early small molecule inhibitors that have demonstrated p47phox-inhibitory effects by disrupting p47phox–p22phox interactions. Celastrol is a triterpenoid compound that yields significant preclinical efficacy in multiple disease models, including AD, RA, cancer, metabolic disorders, and vascular remodeling disorders, among others [138,139,140,141,142,143]. Despite these indications of therapeutic activity, there are concerns about its long-term clinical safety. Celastrol has significant off-target effects including the broad inhibition of heat shock proteins and NF-κB. Treatment with Celastrol also resulted in liver enzyme elevations, and the potential negative effects of chronic Celastrol administration on liver function require further evaluation. Innovative delivery systems may help to improve the therapeutic selectivity and safety of Celastrol [144].
The selenium-based agent Ebselen is another example of a small molecule p47phox inhibitor. Like Celastrol, Ebselen has demonstrated preclinical therapeutic activity for a wide array of indications including cancer, diabetes, neurological disorders, hearing loss, COVID-19 and others [145,146,147,148,149]. Although treatment with Ebselen clearly decreases oxidative stress, it appears to function as a general antioxidant with limited specificity for p47phox-dependent ROS. Nonetheless, Ebselen (SPI-1005) has advanced into clinical trials for multiple therapeutic indications including bipolar disorder, treatment-resistant depression, Meniere’s disease, patients with cochlear implants, type I and II diabetes, COVID-19, and multiple forms of otoprotection (Table 5). Additionally, preclinical efforts are ongoing to develop optimized Ebselen analogs with fewer off-target effects based on recent structure–activity relationship (SAR) studies [150]. Further investigation is needed to determine whether these new Ebselen derivatives offer significant advantages over the parent compound that would support clinical investigation for the potential treatment of p47phox-linked diseases.
Novel SH3-domain binders such as C6 and C14 represent another class of agents that aim to antagonize p47phox function and have demonstrated promise in initial preclinical studies [151]. However, solubility issues may create challenges that hinder clinical evaluation. IP6 (inositol hexaphosphate) also showed potential activity as an inhibitor of p47phox based on its ability to disrupt membrane anchoring in an initial study [152]. However, its poor drug-like properties require optimization for further development. LMH001 is another small molecule inhibitor that has been reported to have efficacy in preclinical models of oxidative stress-related cardiovascular disease [153,154]. More recently, optimized bivalent inhibitors of p47phox-p22phox have been developed with sub-micromolar binding affinities and cellular activity. For example, compound 33 exhibited a Ki50 of 0.24 µM [155]. This impressive initial study provides strong support for further investigation.

5.2. p47phox Agonists

Augmenting p47phox activity to enhance NOX2-mediated ROS production represents a promising therapeutic strategy for conditions characterized by impaired immune responses, such as CGD, chronic infections, and certain cancers. By amplifying the oxidative burst in immune cells, p47phox agonists have the potential to bolster antimicrobial and tumoricidal activities [156]. This strategy could also be very effective for the selective killing of malignant cells that are under constitutive oxidative stress by pushing them over a threshold that triggers ROS-mediated apoptosis [157]. However, the systemic risks of excessive ROS production, including tissue damage and inflammation, necessitate precise control over agonist specificity, delivery, and dosing. Advances in molecular targeting and delivery systems are beginning to address these challenges and may offer a pathway toward safe and effective implementation.
PKC activators like PMA are potent inducers of p47phox phosphorylation at multiple serine residues (Ser303–Ser379) within its autoinhibitory domain [158]. This PTM disrupts intramolecular constraints, facilitates p47phox translocation to the membrane, and stimulates superoxide production in neutrophils. However, there are significant safety issues associated with systemic PMA administration including induction of pro-inflammatory IL-1β and liver toxicity [159]. To mitigate these effects, hydrogels and nanoparticle-encapsulated PMA formulations have been developed for targeted delivery [160]. These innovative delivery strategies are likely to be pivotal in harnessing PKC activators for clinical use.
Toll-like receptor ligands like the TLR7/8 agonist CL097 offer another mechanism to enhance p47phox function by stabilizing phosphorylation at Ser345 through the peptidyl-prolyl isomerase Pin1 [161]. Studies also suggest that immune stimulants may enhance ROS production in CGD neutrophils. This could potentially improve microbial clearance, but efficacy may vary [162]. This mechanism may also enhance immune responses in cancer. However, the effects of TLR agonists on p47phox/NOX2 are not specific. It is clear that other p47phox-independent mechanisms are contributing to their therapeutic activity in different contexts. Structure-based refinement paired with optimized delivery systems will empower the development of more selective agents. Synthetic analogs of the bacterial peptide fMLF provide another strategy to stimulate p47phox activity [163]. They engage formyl peptide receptors on phagocytes to improve residual NOX2 function in CGD neutrophils, increasing ROS production and microbial clearance. The ability of fMLF analogs to selectively activate phagocytes makes them versatile tools. However, their clinical translation requires further optimization to minimize cardiovascular risks.
Beyond these established agonists, novel approaches are expanding the therapeutic landscape. While no direct and specific small molecule activators of p47phox or NOX2 have been reported in the literature to date, this represents an emerging and intriguing area of investigation. Innovative, small molecule-based strategies that leverage targeted activation of p47phox to boost NOX2 activity could be used to enhance immune responses in cancer or directly kill ROS-addicted malignant cells. Other recent efforts have centered around gene therapy and RNA-based approaches. In preclinical studies, lentiviral delivery of functional NCF1 restored ROS in CGD patient neutrophils and reduced infection rates and granuloma size [164]. This approach (PM359) recently advanced into a phase I clinical trial for patients with CGD (Table 5). It will be very interesting to see whether the impressive effects of NCF1 gene restoration that were observed in the preclinical setting can be safely recapitulated in patients. If so, this would represent a groundbreaking development in the field of CGD management. Regardless of the specific modality, it is clear that the future of p47phox agonists lies in precision and personalization. Advances in nanoparticle-based delivery, optogenetics, and gene editing offer unprecedented control over ROS production. This will empower the development of therapies tailored to specific diseases and individual patient profiles. By navigating the delicate interplay between efficacy and safety, p47phox agonists hold the potential to transform the management of immune-deficient and neoplastic diseases.

6. Current Challenges and Future Directions

p47phox functions as a linchpin of NOX2-mediated ROS production. Its precision therapeutic targeting holds transformative potential for diseases characterized by redox dysregulation including CGD, cancer, diabetes and inflammatory disorders. However, its structural complexity, functional pleiotropy, and context-dependent roles present formidable challenges that demand innovative solutions. Overcoming these hurdles to unlock p47phox’s full therapeutic and diagnostic potential will require a convergence of advanced structural biology, precision medicine, and interdisciplinary collaboration.
The dynamic architecture of p47phox is governed by phosphorylation-induced conformational shifts and lipid interactions, which poses a significant barrier to drug design. Its autoinhibitory domain, SH3 domains, and PX domain orchestrate a tightly regulated activation process. This renders traditional small molecule inhibitors or agonists difficult to develop [165]. Moreover, p47phox shares structural homology with the regulatory subunits of other NOX isoforms (e.g., NOX1, NOX4). This complicates efforts to achieve isoform-specific modulation without off-target effects [166]. Recent advances in structural biology are beginning to illuminate druggable allosteric sites within p47phox’s flexible domains [28]. For instance, high-throughput screening coupled with fragment-based drug discovery has identified small molecules that disrupt p47phox’s protein–protein interactions with p22phox or Rac. However, specificity remains elusive due to overlapping binding interfaces [128]. Additionally, p47phox’s non-canonical role in actin remodeling via the Wiskott–Aldrich syndrome protein family verprolin homologous protein (WAVE) complex introduces further complexity as modulation may inadvertently affect immune cell migration or tumor metastasis [167]. These challenges underscore the need for next-generation structural tools like time-resolved crystallography and molecular dynamics simulations to map p47phox’s conformational landscape and design exquisitely selective therapeutics.
The context-dependent nature of p47phox’s function necessitates precise modulation tailored to specific disease states. Achieving this requires a granular understanding of p47phox’s phosphorylation states, expression patterns, and interactome across tissues and cell types. Emerging technologies including single-cell proteomics, spatial transcriptomics, and patient-derived organoid models are poised to address this need by mapping p47phox’s molecular signatures with unprecedented resolution [168]. These insights pave the way for precision therapeutics like proteolysis-targeting chimeras (PROTACs) to selectively degrade hyperactive p47phox in inflammatory contexts or covalent inhibitors that target reactive cysteine residues. Bispecific molecules offer another promising avenue to enhance specificity in cancer immunotherapy.
Therapeutic strategies targeting p47phox are advancing rapidly, propelled by innovations in drug delivery systems and gene editing technologies. Nanoparticle-based platforms like liposomes and hydrogels have been developed to facilitate localized delivery of p47phox modulators to minimize systemic toxicity. For instance, liposome-integrated hydrogel systems have shown promise for cancer treatment and tissue regeneration by enabling controlled release of therapeutic agents [169]. Optogenetic approaches may offer spatiotemporal precision in modulating p47phox activity. Studies have demonstrated that light-activated immune cells can enhance antitumor responses without inducing systemic oxidative damage. This could offer an approach for precise immune modulation [170].
Gene editing techniques, as exemplified by CRISPR-Cas9-mediated correction of the NCF1 gene, present transformative avenues for restoring ROS production in CGD. Targeted repair strategies have successfully restored NOX function in patient-derived cells. This approach offers a potentially curative treatment that has recently advanced into Phase I clinical testing (PM359, Table 5) [171]. Combination therapies are also gaining traction. Evidence suggests that pairing p47phox inhibitors with immune checkpoint inhibitors may reduce tumor progression in murine models by simultaneously attenuating tumor-promoting ROS and enhancing T-cell activity [172]. Furthermore, exploration of the epigenetic regulation of the NCF1 promoter may offer another therapeutic path. DNA hypomethylating agents that have been employed to therapeutically modulate gene expression in cancer cells could potentially be repurposed to reduce p47phox activity through epigenetic modifications in other diseases [173,174].
Structural biologists are currently leveraging cryo-EM and AI-driven molecular modeling to elucidate novel allosteric sites within p47phox and the NOX2 complex. For instance, recent cryo-EM studies have resolved portions of p47phox in active NOX complexes. This provides insights into its interactions and potential regulatory sites. Pharmacologists can utilize this information to design bispecific molecules and covalent inhibitors with enhanced specificity. Notably, the development of bivalent small molecule inhibitors targeting the p47phox–p22phox interaction has already shown promise in selectively inhibiting NOX2 activity [155].
Clinicians can use recent advancements in the understanding of p47phox pathobiology to design patient-specific trials informed by redox profiling and NCF1 genotyping. Such stratification could enhance the efficacy of redox-based therapies by identifying patients who would benefit most from p47phox-targeted interventions. In parallel, the diagnostic and prognostic potential of p47phox is gaining attention, particularly in disorders like rheumatoid arthritis and CGD where altered expression or function of p47phox correlates with disease activity, treatment response, or risk of infection. Advancements in biomarker detection including the use of redox-related biomarkers also support personalized approaches to managing oxidative stress-related diseases. Moreover, exploration of the epigenetic regulation of p47phox should extend beyond DNA methylation to include non-coding RNAs (ncRNAs). Emerging evidence indicates that specific microRNAs can modulate NCF1 expression, which offers a platform for precision modulation of p47phox activity.
In conclusion, p47phox has emerged as a pivotal regulator of redox biology. Its precision targeting offers significant opportunities to address diseases driven by oxidative dysregulation, including immunodeficiency, inflammation, metabolic disease, and cancer. Its structural complexity and context-dependent functions underscore the need for innovative technologies and cross-disciplinary collaboration to harness its therapeutic potential. Emerging research is also exploring p47phox as a potential diagnostic biomarker, although further validation is required. The integration of structural biology, precision therapeutics, and systems-level analysis positions p47phox as a promising target in redox medicine. Continued efforts to overcome current challenges could refine therapeutic strategies and advance the era of precision redox biology.

Author Contributions

Literature review, manuscript preparation: M.E.G.; conceptualization, literature review, manuscript writing and editing: S.S., M.J.C.E., N.L.H., C.M.E., F.B., K.R.K., and W.W.; conceptualization, literature review, figure and table design, manuscript writing and editing: S.T.N., J.S.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by Department of Defense grant W81XWH-22-1-0319 and National Cancer Institute grants R01CA268383, T32CA009213, and P30CA023074. S.S. is a recipient of the University of Arizona Cancer Center ICS-CRTEC fellowship.

Data Availability Statement

As this is a review article, no new data were created. Data availability is not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
NOX2NADPH oxidase 2
CGDChronic granulomatous disease
ROSReactive oxygen species
NOXNADPH oxidase
PRRProline-rich region
O2Superoxide radical
H2O2Hydrogen peroxide
OH•Hydroxyl radical
HOClHypochlorous acid
NCF1Neutrophil cytosolic factor 1
PXPhox homology domain
SH3Src homology domain 3
PKCProtein kinase C
NMRNuclear magnetic resonance
Cryo-EMCryo-electron microscopy
PTMPost-translational modification
UPSUbiquitin proteasome system
PAMPPathogen-associated molecular pattern
DAMPDamage-associated molecular pattern
AIRAutoinhibitory region
FRETFluorescence resonance energy transfer
BRETBioluminescence resonance energy transfer
PPIProtein–protein interaction
TNF-αTumor necrosis factor alpha
PKAProtein kinase A
CK2Casein kinase 2
RARheumatoid arthritis
ADAlzheimer’s disease
PDParkinson’s disease
IBDInflammatory bowel disease
VEGFVascular endothelial growth factor
MMPMatrix metalloproteinases
TGF-βTransforming growth factor beta
NF-κBNuclear factor kappa B
NAFLDNon-alcoholic fatty liver disease
PRRPattern recognition receptor
SARStructure–activity relationship
IP6Inositol hexaphosphate
WAVEWiskott–Aldrich syndrome protein family verprolin homologous protein
PROTACProteolysis targeting chimera

References

  1. Babior, B.M. NADPH oxidase: An update. Blood 1999, 93, 1464–1476. [Google Scholar] [CrossRef] [PubMed]
  2. Quinn, M.T.; Gauss, K.A. Structure and regulation of the neutrophil respiratory burst oxidase: Comparison with nonphagocyte oxidases. J. Leukoc. Biol. 2004, 76, 760–781. [Google Scholar] [CrossRef]
  3. Leto, T.L.; Geiszt, M. Role of Nox family NADPH oxidases in host defense. Antioxid. Redox Signal. 2006, 8, 1549–1561. [Google Scholar] [CrossRef]
  4. Cross, A.R.; Segal, A.W. The NADPH oxidase of professional phagocytes--prototype of the NOX electron transport chain systems. Biochim. Biophys. Acta 2004, 1657, 1–22. [Google Scholar] [CrossRef] [PubMed]
  5. Finkel, T. Signal transduction by reactive oxygen species. J. Cell Biol. 2011, 194, 7–15. [Google Scholar] [CrossRef] [PubMed]
  6. Lambeth, J.D. NOX enzymes and the biology of reactive oxygen. Nat. Rev. Immunol. 2004, 4, 181–189. [Google Scholar] [CrossRef]
  7. El-Benna, J.; Dang, P.M.; Gougerot-Pocidalo, M.A.; Marie, J.C.; Braut-Boucher, F. p47phox, the phagocyte NADPH oxidase/NOX2 organizer: Structure, phosphorylation and implication in diseases. Exp. Mol. Med. 2009, 41, 217–225. [Google Scholar] [CrossRef]
  8. DeLeo, F.R.; Quinn, M.T. Assembly of the phagocyte NADPH oxidase: Molecular interaction of oxidase proteins. J. Leukoc. Biol. 1996, 60, 677–691. [Google Scholar] [CrossRef]
  9. Bedard, K.; Krause, K.H. The NOX family of ROS-generating NADPH oxidases: Physiology and pathophysiology. Physiol. Rev. 2007, 87, 245–313. [Google Scholar] [CrossRef]
  10. Sumimoto, H. Structure, regulation and evolution of Nox-family NADPH oxidases that produce reactive oxygen species. FEBS J. 2008, 275, 3249–3277. [Google Scholar] [CrossRef]
  11. Groemping, Y.; Rittinger, K. Activation and assembly of the NADPH oxidase: A structural perspective. Biochem. J. 2005, 386, 401–416. [Google Scholar] [CrossRef] [PubMed]
  12. Nocella, C.; D’Amico, A.; Cammisotto, V.; Bartimoccia, S.; Castellani, V.; Loffredo, L.; Marini, L.; Ferrara, G.; Testa, M.; Motta, G.; et al. Structure, Activation, and Regulation of NOX2: At the Crossroad between the Innate Immunity and Oxidative Stress-Mediated Pathologies. Antioxidants 2023, 12, 429. [Google Scholar] [CrossRef] [PubMed]
  13. Winkelstein, J.A.; Marino, M.C.; Johnston, R.B., Jr.; Boyle, J.; Curnutte, J.; Gallin, J.I.; Malech, H.L.; Holland, S.M.; Ochs, H.; Quie, P.; et al. Chronic granulomatous disease. Report on a national registry of 368 patients. Medicine 2000, 79, 155–169. [Google Scholar] [CrossRef]
  14. Segal, B.H.; Veys, P.; Malech, H.; Cowan, M.J. Chronic granulomatous disease: Lessons from a rare disorder. Biol. Blood Marrow Transplant. 2011, 17, S123–S131. [Google Scholar] [CrossRef] [PubMed]
  15. Roos, D.; van Bruggen, R.; Meischl, C. Oxidative killing of microbes by neutrophils. Microbes Infect. 2003, 5, 1307–1315. [Google Scholar] [CrossRef]
  16. Clark, R.A.; Volpp, B.D.; Leidal, K.G.; Nauseef, W.M. Two cytosolic components of the human neutrophil respiratory burst oxidase translocate to the plasma membrane during cell activation. J. Clin. Investig. 1990, 85, 714–721. [Google Scholar] [CrossRef]
  17. Kanai, F.; Liu, H.; Field, S.J.; Akbary, H.; Matsuo, T.; Brown, G.E.; Cantley, L.C.; Yaffe, M.B. The PX domains of p47phox and p40phox bind to lipid products of PI(3)K. Nat. Cell Biol. 2001, 3, 675–678. [Google Scholar] [CrossRef]
  18. Stahelin, R.V.; Burian, A.; Bruzik, K.S.; Murray, D.; Cho, W. Membrane binding mechanisms of the PX domains of NADPH oxidase p40phox and p47phox. J. Biol. Chem. 2003, 278, 14469–14479. [Google Scholar] [CrossRef]
  19. Karathanassis, D.; Stahelin, R.V.; Bravo, J.; Perisic, O.; Pacold, C.M.; Cho, W.; Williams, R.L. Binding of the PX domain of p47(phox) to phosphatidylinositol 3,4-bisphosphate and phosphatidic acid is masked by an intramolecular interaction. EMBO J. 2002, 21, 5057–5068. [Google Scholar] [CrossRef]
  20. Sumimoto, H.; Kage, Y.; Nunoi, H.; Sasaki, H.; Nose, T.; Fukumaki, Y.; Ohno, M.; Minakami, S.; Takeshige, K. Role of Src homology 3 domains in assembly and activation of the phagocyte NADPH oxidase. Proc. Natl. Acad. Sci. USA 1994, 91, 5345–5349. [Google Scholar] [CrossRef]
  21. Leto, T.L.; Adams, A.G.; de Mendez, I. Assembly of the phagocyte NADPH oxidase: Binding of Src homology 3 domains to proline-rich targets. Proc. Natl. Acad. Sci. USA 1994, 91, 10650–10654. [Google Scholar] [CrossRef]
  22. Fuchs, A.; Dagher, M.C.; Vignais, P.V. Mapping the domains of interaction of p40phox with both p47phox and p67phox of the neutrophil oxidase complex using the two-hybrid system. J. Biol. Chem. 1995, 270, 5695–5697. [Google Scholar] [CrossRef]
  23. Yuzawa, S.; Suzuki, N.N.; Fujioka, Y.; Ogura, K.; Sumimoto, H.; Inagaki, F. A molecular mechanism for autoinhibition of the tandem SH3 domains of p47phox, the regulatory subunit of the phagocyte NADPH oxidase. Genes Cells 2004, 9, 443–456. [Google Scholar] [CrossRef]
  24. Groemping, Y.; Lapouge, K.; Smerdon, S.J.; Rittinger, K. Molecular basis of phosphorylation-induced activation of the NADPH oxidase. Cell 2003, 113, 343–355. [Google Scholar] [CrossRef] [PubMed]
  25. Han, C.H.; Freeman, J.L.; Lee, T.; Motalebi, S.A.; Lambeth, J.D. Regulation of the neutrophil respiratory burst oxidase. Identification of an activation domain in p67(phox). J. Biol. Chem. 1998, 273, 16663–16668. [Google Scholar] [CrossRef] [PubMed]
  26. Koga, H.; Terasawa, H.; Nunoi, H.; Takeshige, K.; Inagaki, F.; Sumimoto, H. Tetratricopeptide repeat (TPR) motifs of p67(phox) participate in interaction with the small GTPase Rac and activation of the phagocyte NADPH oxidase. J. Biol. Chem. 1999, 274, 25051–25060. [Google Scholar] [CrossRef] [PubMed]
  27. Ziegler, C.S.; Bouchab, L.; Tramier, M.; Durand, D.; Fieschi, F.; Dupre-Crochet, S.; Merola, F.; Nusse, O.; Erard, M. Quantitative live-cell imaging and 3D modeling reveal critical functional features in the cytosolic complex of phagocyte NADPH oxidase. J. Biol. Chem. 2019, 294, 3824–3836. [Google Scholar] [CrossRef]
  28. Aimeur, S.; Fas, B.A.; Serfaty, X.; Santuz, H.; Sacquin-Mora, S.; Bizouarn, T.; Taly, A.; Baciou, L. Structural profiles of the full phagocyte NADPH oxidase unveiled by combining computational biology and experimental knowledge. J. Biol. Chem. 2024, 300, 107943. [Google Scholar] [CrossRef]
  29. Liu, X.; Shi, Y.; Liu, R.; Song, K.; Chen, L. Structure of human phagocyte NADPH oxidase in the activated state. Nature 2024, 627, 189–195. [Google Scholar] [CrossRef]
  30. Johnson, J.L.; Park, J.W.; Benna, J.E.; Faust, L.P.; Inanami, O.; Babior, B.M. Activation of p47(PHOX), a cytosolic subunit of the leukocyte NADPH oxidase. Phosphorylation of ser-359 or ser-370 precedes phosphorylation at other sites and is required for activity. J. Biol. Chem. 1998, 273, 35147–35152. [Google Scholar] [CrossRef]
  31. El-Benna, J.; Dang, P.M.; Gougerot-Pocidalo, M.A. Priming of the neutrophil NADPH oxidase activation: Role of p47phox phosphorylation and NOX2 mobilization to the plasma membrane. Semin. Immunopathol. 2008, 30, 279–289. [Google Scholar] [CrossRef] [PubMed]
  32. Fontayne, A.; Dang, P.M.; Gougerot-Pocidalo, M.A.; El-Benna, J. Phosphorylation of p47phox sites by PKC alpha, beta II, delta, and zeta: Effect on binding to p22phox and on NADPH oxidase activation. Biochemistry 2002, 41, 7743–7750. [Google Scholar] [CrossRef]
  33. Dang, P.M.; Stensballe, A.; Boussetta, T.; Raad, H.; Dewas, C.; Kroviarski, Y.; Hayem, G.; Jensen, O.N.; Gougerot-Pocidalo, M.A.; El-Benna, J. A specific p47phox-serine phosphorylated by convergent MAPKs mediates neutrophil NADPH oxidase priming at inflammatory sites. J. Clin. Investig. 2006, 116, 2033–2043. [Google Scholar] [CrossRef]
  34. Hoyal, C.R.; Gutierrez, A.; Young, B.M.; Catz, S.D.; Lin, J.H.; Tsichlis, P.N.; Babior, B.M. Modulation of p47PHOX activity by site-specific phosphorylation: Akt-dependent activation of the NADPH oxidase. Proc. Natl. Acad. Sci. USA 2003, 100, 5130–5135. [Google Scholar] [CrossRef] [PubMed]
  35. Bengis-Garber, C.; Gruener, N. Protein kinase A downregulates the phosphorylation of p47 phox in human neutrophils: A possible pathway for inhibition of the respiratory burst. Cell Signal. 1996, 8, 291–296. [Google Scholar] [CrossRef] [PubMed]
  36. Raad, H.; Mouawia, H.; Hassan, H.; El-Seblani, M.; Arabi-Derkawi, R.; Boussetta, T.; Gougerot-Pocidalo, M.A.; Dang, P.M.; El-Benna, J. The protein kinase A negatively regulates reactive oxygen species production by phosphorylating gp91phox/NOX2 in human neutrophils. Free Radic. Biol. Med. 2020, 160, 19–27. [Google Scholar] [CrossRef]
  37. Park, H.S.; Lee, S.M.; Lee, J.H.; Kim, Y.S.; Bae, Y.S.; Park, J.W. Phosphorylation of the leucocyte NADPH oxidase subunit p47(phox) by casein kinase 2: Conformation-dependent phosphorylation and modulation of oxidase activity. Biochem. J. 2001, 358, 783–790. [Google Scholar] [CrossRef]
  38. Valenta, H.; Dupre-Crochet, S.; Abdesselem, M.; Bizouarn, T.; Baciou, L.; Nusse, O.; Deniset-Besseau, A.; Erard, M. Consequences of the constitutive NOX2 activity in living cells: Cytosol acidification, apoptosis, and localized lipid peroxidation. Biochim. Biophys. Acta Mol. Cell Res. 2022, 1869, 119276. [Google Scholar] [CrossRef]
  39. Jaquet, V.; Scapozza, L.; Clark, R.A.; Krause, K.H.; Lambeth, J.D. Small-molecule NOX inhibitors: ROS-generating NADPH oxidases as therapeutic targets. Antioxid. Redox Signal. 2009, 11, 2535–2552. [Google Scholar] [CrossRef]
  40. Cipriano, A.; Viviano, M.; Feoli, A.; Milite, C.; Sarno, G.; Castellano, S.; Sbardella, G. NADPH Oxidases: From Molecular Mechanisms to Current Inhibitors. J. Med. Chem. 2023, 66, 11632–11655. [Google Scholar] [CrossRef]
  41. Boussetta, T.; Raad, H.; Bedouhene, S.; Arabi Derkawi, R.; Gougerot-Pocidalo, M.A.; Hayem, G.; Dang, P.M.; El-Benna, J. The peptidyl-prolyl isomerase Pin1 controls GM-CSF-induced priming of NADPH oxidase in human neutrophils and priming at inflammatory sites. Int. Immunopharmacol. 2024, 137, 112425. [Google Scholar] [CrossRef] [PubMed]
  42. Albadrani, M.; Seth, R.K.; Sarkar, S.; Kimono, D.; Mondal, A.; Bose, D.; Porter, D.E.; Scott, G.I.; Brooks, B.; Raychoudhury, S.; et al. Exogenous PP2A inhibitor exacerbates the progression of nonalcoholic fatty liver disease via NOX2-dependent activation of miR21. Am. J. Physiol. Gastrointest. Liver Physiol. 2019, 317, G408–G428. [Google Scholar] [CrossRef] [PubMed]
  43. Vucic, D.; Dixit, V.M.; Wertz, I.E. Ubiquitylation in apoptosis: A post-translational modification at the edge of life and death. Nat. Rev. Mol. Cell Biol. 2011, 12, 439–452. [Google Scholar] [CrossRef]
  44. Choudhary, C.; Kumar, C.; Gnad, F.; Nielsen, M.L.; Rehman, M.; Walther, T.C.; Olsen, J.V.; Mann, M. Lysine acetylation targets protein complexes and co-regulates major cellular functions. Science 2009, 325, 834–840. [Google Scholar] [CrossRef]
  45. Dalle-Donne, I.; Rossi, R.; Giustarini, D.; Colombo, R.; Milzani, A. S-glutathionylation in protein redox regulation. Free Radic. Biol. Med. 2007, 43, 883–898. [Google Scholar] [CrossRef]
  46. Foster, M.W.; McMahon, T.J.; Stamler, J.S. S-nitrosylation in health and disease. Trends Mol. Med. 2003, 9, 160–168. [Google Scholar] [CrossRef]
  47. Hart, G.W.; Slawson, C.; Ramirez-Correa, G.; Lagerlof, O. Cross talk between O-GlcNAcylation and phosphorylation: Roles in signaling, transcription, and chronic disease. Annu. Rev. Biochem. 2011, 80, 825–858. [Google Scholar] [CrossRef]
  48. Belambri, S.A.; Marzaioli, V.; Hurtado-Nedelec, M.; Pintard, C.; Liang, S.; Liu, Y.; Boussetta, T.; Gougerot-Pocidalo, M.A.; Ye, R.D.; Dang, P.M.; et al. Impaired p47phox phosphorylation in neutrophils from patients with p67phox-deficient chronic granulomatous disease. Blood 2022, 139, 2512–2522. [Google Scholar] [CrossRef] [PubMed]
  49. Roesler, J.; Curnutte, J.T.; Rae, J.; Barrett, D.; Patino, P.; Chanock, S.J.; Goerlach, A. Recombination events between the p47-phox gene and its highly homologous pseudogenes are the main cause of autosomal recessive chronic granulomatous disease. Blood 2000, 95, 2150–2156. [Google Scholar] [CrossRef]
  50. Jackson, S.H.; Gallin, J.I.; Holland, S.M. The p47phox mouse knock-out model of chronic granulomatous disease. J. Exp. Med. 1995, 182, 751–758. [Google Scholar] [CrossRef]
  51. Morgenstern, D.E.; Gifford, M.A.; Li, L.L.; Doerschuk, C.M.; Dinauer, M.C. Absence of respiratory burst in X-linked chronic granulomatous disease mice leads to abnormalities in both host defense and inflammatory response to Aspergillus fumigatus. J. Exp. Med. 1997, 185, 207–218. [Google Scholar] [CrossRef] [PubMed]
  52. Kang, S.W.; Chae, H.Z.; Seo, M.S.; Kim, K.; Baines, I.C.; Rhee, S.G. Mammalian peroxiredoxin isoforms can reduce hydrogen peroxide generated in response to growth factors and tumor necrosis factor-alpha. J. Biol. Chem. 1998, 273, 6297–6302. [Google Scholar] [CrossRef]
  53. Bagaitkar, J.; Huang, J.; Zeng, M.Y.; Pech, N.K.; Monlish, D.A.; Perez-Zapata, L.J.; Miralda, I.; Schuettpelz, L.G.; Dinauer, M.C. NADPH oxidase activation regulates apoptotic neutrophil clearance by murine macrophages. Blood 2018, 131, 2367–2378. [Google Scholar] [CrossRef]
  54. Pollock, J.D.; Williams, D.A.; Gifford, M.A.; Li, L.L.; Du, X.; Fisherman, J.; Orkin, S.H.; Doerschuk, C.M.; Dinauer, M.C. Mouse model of X-linked chronic granulomatous disease, an inherited defect in phagocyte superoxide production. Nat. Genet. 1995, 9, 202–209. [Google Scholar] [CrossRef]
  55. Darbinian, E.; Mlaga, K.D.; Chandrasekaran, P.; Han, Y.; Donko, A.; Desjardins, A.; Leto, T.L.; Holland, S.M.; Poudrier, J.; Falcone, E.L. Genotype-specific immune responses at the intestinal barrier predispose to colitis in chronic granulomatous disease in mice. Blood 2025, 145, 3153–3165. [Google Scholar] [CrossRef]
  56. Ott, M.G.; Schmidt, M.; Schwarzwaelder, K.; Stein, S.; Siler, U.; Koehl, U.; Glimm, H.; Kuhlcke, K.; Schilz, A.; Kunkel, H.; et al. Correction of X-linked chronic granulomatous disease by gene therapy, augmented by insertional activation of MDS1-EVI1, PRDM16 or SETBP1. Nat. Med. 2006, 12, 401–409. [Google Scholar] [CrossRef] [PubMed]
  57. Hacein-Bey-Abina, S.; Garrigue, A.; Wang, G.P.; Soulier, J.; Lim, A.; Morillon, E.; Clappier, E.; Caccavelli, L.; Delabesse, E.; Beldjord, K.; et al. Insertional oncogenesis in 4 patients after retrovirus-mediated gene therapy of SCID-X1. J. Clin. Investig. 2008, 118, 3132–3142. [Google Scholar] [CrossRef]
  58. Fuentes, E.; Gibbins, J.M.; Holbrook, L.M.; Palomo, I. NADPH oxidase 2 (NOX2): A key target of oxidative stress-mediated platelet activation and thrombosis. Trends Cardiovasc. Med. 2018, 28, 429–434. [Google Scholar] [CrossRef] [PubMed]
  59. Griendling, K.K.; Sorescu, D.; Ushio-Fukai, M. NAD(P)H oxidase: Role in cardiovascular biology and disease. Circ. Res. 2000, 86, 494–501. [Google Scholar] [CrossRef]
  60. Judkins, C.P.; Diep, H.; Broughton, B.R.; Mast, A.E.; Hooker, E.U.; Miller, A.A.; Selemidis, S.; Dusting, G.J.; Sobey, C.G.; Drummond, G.R. Direct evidence of a role for Nox2 in superoxide production, reduced nitric oxide bioavailability, and early atherosclerotic plaque formation in ApoE−/− mice. Am. J. Physiol. Heart Circ. Physiol. 2010, 298, H24–H32. [Google Scholar] [CrossRef]
  61. Sorescu, D.; Weiss, D.; Lassegue, B.; Clempus, R.E.; Szocs, K.; Sorescu, G.P.; Valppu, L.; Quinn, M.T.; Lambeth, J.D.; Vega, J.D.; et al. Superoxide production and expression of nox family proteins in human atherosclerosis. Circulation 2002, 105, 1429–1435. [Google Scholar] [CrossRef]
  62. Touyz, R.M.; Schiffrin, E.L. Increased generation of superoxide by angiotensin II in smooth muscle cells from resistance arteries of hypertensive patients: Role of phospholipase D-dependent NAD(P)H oxidase-sensitive pathways. J. Hypertens. 2001, 19, 1245–1254. [Google Scholar] [CrossRef] [PubMed]
  63. Lassegue, B.; Sorescu, D.; Szocs, K.; Yin, Q.; Akers, M.; Zhang, Y.; Grant, S.L.; Lambeth, J.D.; Griendling, K.K. Novel gp91(phox) homologues in vascular smooth muscle cells: Nox1 mediates angiotensin II-induced superoxide formation and redox-sensitive signaling pathways. Circ. Res. 2001, 88, 888–894. [Google Scholar] [CrossRef]
  64. Clempus, R.E.; Griendling, K.K. Reactive oxygen species signaling in vascular smooth muscle cells. Cardiovasc. Res. 2006, 71, 216–225. [Google Scholar] [CrossRef]
  65. Lu, W.; Lin, Y.; Haider, N.; Moly, P.; Wang, L.; Zhou, W. Ginsenoside Rb1 protects human vascular smooth muscle cells against resistin-induced oxidative stress and dysfunction. Front. Cardiovasc. Med. 2023, 10, 1164547. [Google Scholar] [CrossRef]
  66. Manea, S.A.; Vlad, M.L.; Lazar, A.G.; Muresian, H.; Simionescu, M.; Manea, A. SET7 lysine methyltransferase mediates the up-regulation of NADPH oxidase expression, oxidative stress, and NLRP3 inflammasome priming in atherosclerosis. J. Transl. Med. 2025, 23, 339. [Google Scholar] [CrossRef] [PubMed]
  67. Harrison, C.B.; Drummond, G.R.; Sobey, C.G.; Selemidis, S. Evidence that nitric oxide inhibits vascular inflammation and superoxide production via a p47phox-dependent mechanism in mice. Clin. Exp. Pharmacol. Physiol. 2010, 37, 429–434. [Google Scholar] [CrossRef] [PubMed]
  68. Hsich, E.; Segal, B.H.; Pagano, P.J.; Rey, F.E.; Paigen, B.; Deleonardis, J.; Hoyt, R.F.; Holland, S.M.; Finkel, T. Vascular effects following homozygous disruption of p47(phox): An essential component of NADPH oxidase. Circulation 2000, 101, 1234–1236. [Google Scholar] [CrossRef]
  69. Guzik, T.J.; Sadowski, J.; Guzik, B.; Jopek, A.; Kapelak, B.; Przybylowski, P.; Wierzbicki, K.; Korbut, R.; Harrison, D.G.; Channon, K.M. Coronary artery superoxide production and nox isoform expression in human coronary artery disease. Arterioscler. Thromb. Vasc. Biol. 2006, 26, 333–339. [Google Scholar] [CrossRef]
  70. Touyz, R.M.; Schiffrin, E.L. Reactive oxygen species in vascular biology: Implications in hypertension. Histochem. Cell Biol. 2004, 122, 339–352. [Google Scholar] [CrossRef]
  71. Jaquet, V.; Marcoux, J.; Forest, E.; Leidal, K.G.; McCormick, S.; Westermaier, Y.; Perozzo, R.; Plastre, O.; Fioraso-Cartier, L.; Diebold, B.; et al. NADPH oxidase (NOX) isoforms are inhibited by celastrol with a dual mode of action. Br. J. Pharmacol. 2011, 164, 507–520. [Google Scholar] [CrossRef]
  72. Chu, S.; Mao, X.; Guo, H.; Wang, L.; Li, Z.; Zhang, Y.; Wang, Y.; Wang, H.; Zhang, X.; Peng, W. Indoxyl sulfate potentiates endothelial dysfunction via reciprocal role for reactive oxygen species and RhoA/ROCK signaling in 5/6 nephrectomized rats. Free Radic. Res. 2017, 51, 237–252. [Google Scholar] [CrossRef]
  73. Ravichandran, K.A.; Heneka, M.T. Inflammasomes in neurological disorders—Mechanisms and therapeutic potential. Nat. Rev. Neurol. 2024, 20, 67–83. [Google Scholar] [CrossRef] [PubMed]
  74. Qin, L.; Liu, Y.; Wang, T.; Wei, S.J.; Block, M.L.; Wilson, B.; Liu, B.; Hong, J.S. NADPH oxidase mediates lipopolysaccharide-induced neurotoxicity and proinflammatory gene expression in activated microglia. J. Biol. Chem. 2004, 279, 1415–1421. [Google Scholar] [CrossRef]
  75. Block, M.L.; Zecca, L.; Hong, J.S. Microglia-mediated neurotoxicity: Uncovering the molecular mechanisms. Nat. Rev. Neurosci. 2007, 8, 57–69. [Google Scholar] [CrossRef] [PubMed]
  76. Wu, D.C.; Teismann, P.; Tieu, K.; Vila, M.; Jackson-Lewis, V.; Ischiropoulos, H.; Przedborski, S. NADPH oxidase mediates oxidative stress in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine model of Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2003, 100, 6145–6150. [Google Scholar] [CrossRef] [PubMed]
  77. Glass, C.K.; Saijo, K.; Winner, B.; Marchetto, M.C.; Gage, F.H. Mechanisms underlying inflammation in neurodegeneration. Cell 2010, 140, 918–934. [Google Scholar] [CrossRef]
  78. Gong, P.; Chen, Y.Q.; Lin, A.H.; Zhang, H.B.; Zhang, Y.; Ye, R.D.; Yu, Y. p47(phox) deficiency improves cognitive impairment and attenuates tau hyperphosphorylation in mouse models of AD. Alzheimers Res. Ther. 2020, 12, 146. [Google Scholar] [CrossRef]
  79. Simpson, D.S.A.; Oliver, P.L. ROS Generation in Microglia: Understanding Oxidative Stress and Inflammation in Neurodegenerative Disease. Antioxidants 2020, 9, 743. [Google Scholar] [CrossRef]
  80. Birmann, P.T.; Casaril, A.M.; Abenante, L.; Penteado, F.; Bruning, C.A.; Savegnago, L.; Lenardao, E.J. Neuropharmacology of Organoselenium Compounds in Mental Disorders and Degenerative Diseases. Curr. Med. Chem. 2023, 30, 2357–2395. [Google Scholar] [CrossRef]
  81. Dustin, C.M.; Shiva, S.S.; Vazquez, A.; Saeed, A.; Pascoal, T.; Cifuentes-Pagano, E.; Pagano, P.J. NOX2 in Alzheimer’s and Parkinson’s disease. Redox Biol. 2024, 78, 103433. [Google Scholar] [CrossRef] [PubMed]
  82. Kuber, B.; Fadnavis, M.; Chatterjee, B. Role of angiotensin receptor blockers in the context of Alzheimer’s disease. Fundam. Clin. Pharmacol. 2023, 37, 429–445. [Google Scholar] [CrossRef] [PubMed]
  83. Dragoi, C.M.; Diaconu, C.C.; Nicolae, A.C.; Dumitrescu, I.B. Redox Homeostasis and Molecular Biomarkers in Precision Therapy for Cardiovascular Diseases. Antioxidants 2024, 13, 1163. [Google Scholar] [CrossRef]
  84. McInnes, I.B.; Schett, G. The pathogenesis of rheumatoid arthritis. N. Engl. J. Med. 2011, 365, 2205–2219. [Google Scholar] [CrossRef] [PubMed]
  85. Lin, W.; Shen, P.; Song, Y.; Huang, Y.; Tu, S. Reactive Oxygen Species in Autoimmune Cells: Function, Differentiation, and Metabolism. Front. Immunol. 2021, 12, 635021. [Google Scholar] [CrossRef]
  86. Chenevier-Gobeaux, C.; Lemarechal, H.; Bonnefont-Rousselot, D.; Poiraudeau, S.; Ekindjian, O.G.; Borderie, D. Superoxide production and NADPH oxidase expression in human rheumatoid synovial cells: Regulation by interleukin-1beta and tumour necrosis factor-alpha. Inflamm. Res. 2006, 55, 483–490. [Google Scholar] [CrossRef]
  87. Ryo, A.; Suizu, F.; Yoshida, Y.; Perrem, K.; Liou, Y.C.; Wulf, G.; Rottapel, R.; Yamaoka, S.; Lu, K.P. Regulation of NF-kappaB signaling by Pin1-dependent prolyl isomerization and ubiquitin-mediated proteolysis of p65/RelA. Mol. Cell 2003, 12, 1413–1426. [Google Scholar] [CrossRef]
  88. Jacobsson, L.T.; Turesson, C.; Nilsson, J.A.; Petersson, I.F.; Lindqvist, E.; Saxne, T.; Geborek, P. Treatment with TNF blockers and mortality risk in patients with rheumatoid arthritis. Ann. Rheum. Dis. 2007, 66, 670–675. [Google Scholar] [CrossRef]
  89. Ma, Y.; Lin, H.; Li, Y.; An, Z. Amentoflavone Induces Ferroptosis to Alleviate Proliferation, Migration, Invasion and Inflammation in Rheumatoid Arthritis Fibroblast-like Synoviocytes by Inhibiting PIN1. Cell Biochem. Biophys. 2025, 83, 1299–1312. [Google Scholar] [CrossRef]
  90. Roessner, A.; Kuester, D.; Malfertheiner, P.; Schneider-Stock, R. Oxidative stress in ulcerative colitis-associated carcinogenesis. Pathol. Res. Pract. 2008, 204, 511–524. [Google Scholar] [CrossRef]
  91. Falcone, E.L.; Abusleme, L.; Swamydas, M.; Lionakis, M.S.; Ding, L.; Hsu, A.P.; Zelazny, A.M.; Moutsopoulos, N.M.; Kuhns, D.B.; Deming, C.; et al. Colitis susceptibility in p47phox−/− mice is mediated by the microbiome. Microbiome 2016, 4, 13. [Google Scholar] [CrossRef] [PubMed]
  92. Nunoi, H.; Nakamura, H.; Nishimura, T.; Matsukura, M. Recent topics and advanced therapies in chronic granulomatous disease. Hum. Cell 2023, 36, 515–527. [Google Scholar] [CrossRef]
  93. Xavier, R.J.; Podolsky, D.K. Unravelling the pathogenesis of inflammatory bowel disease. Nature 2007, 448, 427–434. [Google Scholar] [CrossRef] [PubMed]
  94. Lupu, A.; Stoleriu, G.; Nedelcu, A.H.; Perju, S.N.; Gavrilovici, C.; Baciu, G.; Mihai, C.M.; Chisnoiu, T.; Morariu, I.D.; Grigore, E.; et al. Overview of Oxidative Stress in Systemic Lupus Erythematosus. Antioxidants 2025, 14, 303. [Google Scholar] [CrossRef] [PubMed]
  95. Lowes, M.A.; Bowcock, A.M.; Krueger, J.G. Pathogenesis and therapy of psoriasis. Nature 2007, 445, 866–873. [Google Scholar] [CrossRef]
  96. Liu, F.C.; Yu, H.P.; Chen, P.J.; Yang, H.W.; Chang, S.H.; Tzeng, C.C.; Cheng, W.J.; Chen, Y.R.; Chen, Y.L.; Hwang, T.L. A novel NOX2 inhibitor attenuates human neutrophil oxidative stress and ameliorates inflammatory arthritis in mice. Redox Biol. 2019, 26, 101273. [Google Scholar] [CrossRef]
  97. Aggarwal, H.; Pathak, P.; Gupta, S.K.; Kumar, Y.; Jagavelu, K.; Dikshit, M. Serum and cecal metabolic profile of the insulin resistant and dyslipidemic p47(phox) knockout mice. Free Radic. Res. 2022, 56, 483–497. [Google Scholar] [CrossRef]
  98. Chen, B.; Li, T.; Wu, Y.; Song, L.; Wang, Y.; Bian, Y.; Qiu, Y.; Yang, Z. Lipotoxicity: A New Perspective in Type 2 Diabetes Mellitus. Diabetes Metab. Syndr. Obes. 2025, 18, 1223–1237. [Google Scholar] [CrossRef]
  99. Laddha, A.P.; Kulkarni, Y.A. NADPH oxidase: A membrane-bound enzyme and its inhibitors in diabetic complications. Eur. J. Pharmacol. 2020, 881, 173206. [Google Scholar] [CrossRef]
  100. Du, X.L.; Edelstein, D.; Rossetti, L.; Fantus, I.G.; Goldberg, H.; Ziyadeh, F.; Wu, J.; Brownlee, M. Hyperglycemia-induced mitochondrial superoxide overproduction activates the hexosamine pathway and induces plasminogen activator inhibitor-1 expression by increasing Sp1 glycosylation. Proc. Natl. Acad. Sci. USA 2000, 97, 12222–12226. [Google Scholar] [CrossRef]
  101. Du, Y.; Veenstra, A.; Palczewski, K.; Kern, T.S. Photoreceptor cells are major contributors to diabetes-induced oxidative stress and local inflammation in the retina. Proc. Natl. Acad. Sci. USA 2013, 110, 16586–16591. [Google Scholar] [CrossRef] [PubMed]
  102. Li, B.; Huo, Y.; Zhang, K.; Chang, L.; Zhang, H.; Wang, X.; Li, L.; Hu, Z. Development and validation of outcome prediction models for acute kidney injury patients undergoing continuous renal replacement therapy. Front. Med. 2022, 9, 853989. [Google Scholar] [CrossRef]
  103. Tojo, A.; Asaba, K.; Onozato, M.L. Suppressing renal NADPH oxidase to treat diabetic nephropathy. Expert Opin. Ther. Targets 2007, 11, 1011–1018. [Google Scholar] [CrossRef] [PubMed]
  104. Liu, G.C.; Fang, F.; Zhou, J.; Koulajian, K.; Yang, S.; Lam, L.; Reich, H.N.; John, R.; Herzenberg, A.M.; Giacca, A.; et al. Deletion of p47phox attenuates the progression of diabetic nephropathy and reduces the severity of diabetes in the Akita mouse. Diabetologia 2012, 55, 2522–2532. [Google Scholar] [CrossRef] [PubMed]
  105. De Minicis, S.; Seki, E.; Paik, Y.H.; Osterreicher, C.H.; Kodama, Y.; Kluwe, J.; Torozzi, L.; Miyai, K.; Benedetti, A.; Schwabe, R.F.; et al. Role and cellular source of nicotinamide adenine dinucleotide phosphate oxidase in hepatic fibrosis. Hepatology 2010, 52, 1420–1430. [Google Scholar] [CrossRef]
  106. Ni, X.; Duan, L.; Bao, Y.; Li, J.; Zhang, X.; Jia, D.; Wu, N. Circ_005077 accelerates myocardial lipotoxicity induced by high-fat diet via CyPA/p47PHOX mediated ferroptosis. Cardiovasc. Diabetol. 2024, 23, 129. [Google Scholar] [CrossRef]
  107. Furukawa, S.; Fujita, T.; Shimabukuro, M.; Iwaki, M.; Yamada, Y.; Nakajima, Y.; Nakayama, O.; Makishima, M.; Matsuda, M.; Shimomura, I. Increased oxidative stress in obesity and its impact on metabolic syndrome. J. Clin. Investig. 2004, 114, 1752–1761. [Google Scholar] [CrossRef]
  108. Ushio-Fukai, M.; Nakamura, Y. Reactive oxygen species and angiogenesis: NADPH oxidase as target for cancer therapy. Cancer Lett. 2008, 266, 37–52. [Google Scholar] [CrossRef]
  109. Weinberg, F.; Hamanaka, R.; Wheaton, W.W.; Weinberg, S.; Joseph, J.; Lopez, M.; Kalyanaraman, B.; Mutlu, G.M.; Budinger, G.R.; Chandel, N.S. Mitochondrial metabolism and ROS generation are essential for Kras-mediated tumorigenicity. Proc. Natl. Acad. Sci. USA 2010, 107, 8788–8793. [Google Scholar] [CrossRef]
  110. Kelkka, T.; Pizzolla, A.; Laurila, J.P.; Friman, T.; Gustafsson, R.; Kallberg, E.; Olsson, O.; Leanderson, T.; Rubin, K.; Salmi, M.; et al. Mice lacking NCF1 exhibit reduced growth of implanted melanoma and carcinoma tumors. PLoS ONE 2013, 8, e84148. [Google Scholar] [CrossRef]
  111. Vuong, H.; Cheng, F.; Lin, C.C.; Zhao, Z. Functional consequences of somatic mutations in cancer using protein pocket-based prioritization approach. Genome Med. 2014, 6, 81. [Google Scholar] [CrossRef]
  112. van der Weyden, L.; Speak, A.O.; Swiatkowska, A.; Clare, S.; Schejtman, A.; Santilli, G.; Arends, M.J.; Adams, D.J. Pulmonary metastatic colonisation and granulomas in NOX2-deficient mice. J. Pathol. 2018, 246, 300–310. [Google Scholar] [CrossRef] [PubMed]
  113. Yager, J.D.; Davidson, N.E. Estrogen carcinogenesis in breast cancer. N. Engl. J. Med. 2006, 354, 270–282. [Google Scholar] [CrossRef]
  114. Roy, K.; Wu, Y.; Meitzler, J.L.; Juhasz, A.; Liu, H.; Jiang, G.; Lu, J.; Antony, S.; Doroshow, J.H. NADPH oxidases and cancer. Clin. Sci. (Lond.) 2015, 128, 863–875. [Google Scholar] [CrossRef]
  115. Zheng, K.; Li, Y.; Wang, S.; Wang, X.; Liao, C.; Hu, X.; Fan, L.; Kang, Q.; Zeng, Y.; Wu, X.; et al. Inhibition of autophagosome-lysosome fusion by ginsenoside Ro via the ESR2-NCF1-ROS pathway sensitizes esophageal cancer cells to 5-fluorouracil-induced cell death via the CHEK1-mediated DNA damage checkpoint. Autophagy 2016, 12, 1593–1613. [Google Scholar] [CrossRef] [PubMed]
  116. Paolillo, R.; Boulanger, M.; Gatel, P.; Gabellier, L.; De Toledo, M.; Tempe, D.; Hallal, R.; Akl, D.; Moreaux, J.; Baik, H.; et al. The NADPH oxidase NOX2 is a marker of adverse prognosis involved in chemoresistance of acute myeloid leukemias. Haematologica 2022, 107, 2562–2575. [Google Scholar] [CrossRef]
  117. Quilaqueo-Millaqueo, N.; Brown-Brown, D.A.; Vidal-Vidal, J.A.; Niechi, I. NOX proteins and ROS generation: Role in invadopodia formation and cancer cell invasion. Biol. Res. 2024, 57, 98. [Google Scholar] [CrossRef] [PubMed]
  118. Vallee, A.; Lecarpentier, Y. Crosstalk Between Peroxisome Proliferator-Activated Receptor Gamma and the Canonical WNT/beta-Catenin Pathway in Chronic Inflammation and Oxidative Stress During Carcinogenesis. Front. Immunol. 2018, 9, 745. [Google Scholar] [CrossRef]
  119. Ye, S.X.Y.; Li, J.; Zheng, Y.; Sun, P.; Yang, W. p47phox contributes to inflammation and tumor promotion in colitis-associated cancer through NF-κB activation. J. Immunol. 2016, 196, 153.129. [Google Scholar]
  120. Dhiman, M.; Garg, N.J. P47phox−/− mice are compromised in expansion and activation of CD8+ T cells and susceptible to Trypanosoma cruzi infection. PLoS Pathog. 2014, 10, e1004516. [Google Scholar] [CrossRef]
  121. Ushio-Fukai, M. Redox signaling in angiogenesis: Role of NADPH oxidase. Cardiovasc. Res. 2006, 71, 226–235. [Google Scholar] [CrossRef]
  122. Chen, Y.; He, F.; Wang, R.; Yao, M.; Li, Y.; Guo, D.; He, S. NCF1/2/4 Are Prognostic Biomarkers Related to the Immune Infiltration of Kidney Renal Clear Cell Carcinoma. Biomed. Res. Int. 2021, 2021, 5954036. [Google Scholar] [CrossRef] [PubMed]
  123. Griess, B.; Mir, S.; Datta, K.; Teoh-Fitzgerald, M. Scavenging reactive oxygen species selectively inhibits M2 macrophage polarization and their pro-tumorigenic function in part, via Stat3 suppression. Free Radic. Biol. Med. 2020, 147, 48–60. [Google Scholar] [CrossRef] [PubMed]
  124. Liu, X.; Pei, C.; Yan, S.; Liu, G.; Liu, G.; Chen, W.; Cui, Y.; Liu, Y. NADPH oxidase 1-dependent ROS is crucial for TLR4 signaling to promote tumor metastasis of non-small cell lung cancer. Tumour Biol. 2015, 36, 1493–1502. [Google Scholar] [CrossRef]
  125. Lee, J.K.; Edderkaoui, M.; Truong, P.; Ohno, I.; Jang, K.T.; Berti, A.; Pandol, S.J.; Gukovskaya, A.S. NADPH oxidase promotes pancreatic cancer cell survival via inhibiting JAK2 dephosphorylation by tyrosine phosphatases. Gastroenterology 2007, 133, 1637–1648. [Google Scholar] [CrossRef] [PubMed]
  126. Furuya-Ikude, C.; Kitta, A.; Tomonobu, N.; Kawasaki, Y.; Sakaguchi, M.; Kondo, E. NCF-1 plays a pivotal role in the survival of adenocarcinoma cells of pancreatic and gastric origins. In Vitro Cell Dev. Biol. Anim. 2024, 60, 1151–1159. [Google Scholar] [CrossRef]
  127. Liu, R.; Peng, L.; Zhou, L.; Huang, Z.; Zhou, C.; Huang, C. Oxidative Stress in Cancer Immunotherapy: Molecular Mechanisms and Potential Applications. Antioxidants 2022, 11, 853. [Google Scholar] [CrossRef]
  128. Solbak, S.M.O.; Zang, J.; Narayanan, D.; Hoj, L.J.; Bucciarelli, S.; Softley, C.; Meier, S.; Langkilde, A.E.; Gotfredsen, C.H.; Sattler, M.; et al. Developing Inhibitors of the p47phox-p22phox Protein-Protein Interaction by Fragment-Based Drug Discovery. J. Med. Chem. 2020, 63, 1156–1177. [Google Scholar] [CrossRef]
  129. Rey, F.E.; Cifuentes, M.E.; Kiarash, A.; Quinn, M.T.; Pagano, P.J. Novel competitive inhibitor of NAD(P)H oxidase assembly attenuates vascular O2 and systolic blood pressure in mice. Circ. Res. 2001, 89, 408–414. [Google Scholar] [CrossRef]
  130. Cifuentes-Pagano, E.; Pagano, P.J. Rational Design and Delivery of NOX-Inhibitory Peptides. Methods Mol. Biol. 2019, 1982, 417–428. [Google Scholar] [CrossRef]
  131. Ebrahimian, T.; Li, M.W.; Lemarie, C.A.; Simeone, S.M.; Pagano, P.J.; Gaestel, M.; Paradis, P.; Wassmann, S.; Schiffrin, E.L. Mitogen-activated protein kinase-activated protein kinase 2 in angiotensin II-induced inflammation and hypertension: Regulation of oxidative stress. Hypertension 2011, 57, 245–254. [Google Scholar] [CrossRef] [PubMed]
  132. DiStasi, M.R.; Mund, J.A.; Bohlen, H.G.; Miller, S.J.; Ingram, D.A.; Dalsing, M.C.; Unthank, J.L. Impaired compensation to femoral artery ligation in diet-induced obese mice is primarily mediated via suppression of collateral growth by Nox2 and p47phox. Am. J. Physiol. Heart Circ. Physiol. 2015, 309, H1207–H1217. [Google Scholar] [CrossRef]
  133. Liu, X.; Lei, Z.; Gilhooly, D.; He, J.; Li, Y.; Ritzel, R.M.; Li, H.; Wu, L.J.; Liu, S.; Wu, J. Traumatic brain injury-induced inflammatory changes in the olfactory bulb disrupt neuronal networks leading to olfactory dysfunction. Brain Behav. Immun. 2023, 114, 22–45. [Google Scholar] [CrossRef]
  134. Sabirzhanov, B.; Li, Y.; Coll-Miro, M.; Matyas, J.J.; He, J.; Kumar, A.; Ward, N.; Yu, J.; Faden, A.I.; Wu, J. Inhibition of NOX2 signaling limits pain-related behavior and improves motor function in male mice after spinal cord injury: Participation of IL-10/miR-155 pathways. Brain Behav. Immun. 2019, 80, 73–87. [Google Scholar] [CrossRef] [PubMed]
  135. Craik, D.J.; Fairlie, D.P.; Liras, S.; Price, D. The future of peptide-based drugs. Chem. Biol. Drug Des. 2013, 81, 136–147. [Google Scholar] [CrossRef]
  136. Cifuentes-Pagano, M.E.; Meijles, D.N.; Pagano, P.J. Nox Inhibitors & Therapies: Rational Design of Peptidic and Small Molecule Inhibitors. Curr. Pharm. Des. 2015, 21, 6023–6035. [Google Scholar] [CrossRef] [PubMed]
  137. Li, Y.; Cifuentes-Pagano, E.; DeVallance, E.R.; de Jesus, D.S.; Sahoo, S.; Meijles, D.N.; Koes, D.; Camacho, C.J.; Ross, M.; St Croix, C.; et al. NADPH oxidase 2 inhibitors CPP11G and CPP11H attenuate endothelial cell inflammation & vessel dysfunction and restore mouse hind-limb flow. Redox Biol. 2019, 22, 101143. [Google Scholar] [CrossRef]
  138. Wang, X.; Abu Bakar, M.H.; Liqun, S.; Kassim, M.A.; Shariff, K.A.; Karunakaran, T. Targeting metabolic diseases with celastrol: A comprehensive review of anti-inflammatory mechanisms and therapeutic potential. J. Ethnopharmacol. 2025, 344, 119560. [Google Scholar] [CrossRef]
  139. Lei, H.; Ruan, Y.; Ding, R.; Li, H.; Zhang, X.; Ji, X.; Wang, Q.; Lv, S. The role of celastrol in inflammation and diseases. Inflamm. Res. 2025, 74, 23. [Google Scholar] [CrossRef]
  140. Zhao, Y.; Miettinen, K.; Kampranis, S.C. Celastrol: A century-long journey from the isolation to the biotechnological production and the development of an antiobesity drug. Curr. Opin. Plant Biol. 2024, 81, 102615. [Google Scholar] [CrossRef]
  141. Tan, J.L.; Yi, J.; Cao, X.Y.; Wang, F.Y.; Xie, S.L.; Zhou, L.L.; Qin, L.; Dai, A.G. Celastrol: The new dawn in the treatment of vascular remodeling diseases. Biomed. Pharmacother. 2023, 158, 114177. [Google Scholar] [CrossRef] [PubMed]
  142. Wang, C.; Dai, S.; Zhao, X.; Zhang, Y.; Gong, L.; Fu, K.; Ma, C.; Peng, C.; Li, Y. Celastrol as an emerging anticancer agent: Current status, challenges and therapeutic strategies. Biomed. Pharmacother. 2023, 163, 114882. [Google Scholar] [CrossRef]
  143. Robertson, R.P. Antioxidants for Early Treatment of Type 2 Diabetes in Rodents and Humans: Lost in Translation? Diabetes 2024, 73, 653–658. [Google Scholar] [CrossRef]
  144. Sun, Y.; Wang, C.; Li, X.; Lu, J.; Wang, M. Recent advances in drug delivery of celastrol for enhancing efficiency and reducing the toxicity. Front. Pharmacol. 2024, 15, 1137289. [Google Scholar] [CrossRef]
  145. Ali, F.; Alom, S.; Ali, S.R.; Kondoli, B.; Sadhu, P.; Borah, C.; Kakoti, B.B.; Ghosh, S.K.; Shakya, A.; Ahmed, A.B.; et al. Ebselen: A Review on its Synthesis, Derivatives, Anticancer Efficacy and Utility in Combating SARS-CoV-2. Mini Rev. Med. Chem. 2024, 24, 1203–1225. [Google Scholar] [CrossRef] [PubMed]
  146. Ramli, F.F.; Cowen, P.J.; Godlewska, B.R. The Potential Use of Ebselen in Treatment-Resistant Depression. Pharmaceuticals 2022, 15, 485. [Google Scholar] [CrossRef] [PubMed]
  147. Santi, C.; Scimmi, C.; Sancineto, L. Ebselen and Analogues: Pharmacological Properties and Synthetic Strategies for Their Preparation. Molecules 2021, 26, 4230. [Google Scholar] [CrossRef] [PubMed]
  148. Kil, J.; Harruff, E.E.; Longenecker, R.J. Development of ebselen for the treatment of sensorineural hearing loss and tinnitus. Hear. Res. 2022, 413, 108209. [Google Scholar] [CrossRef]
  149. Wang, J.; Wang, P.; Dong, C.; Zhao, Y.; Zhou, J.; Yuan, C.; Zou, L. Mechanisms of ebselen as a therapeutic and its pharmacology applications. Future Med. Chem. 2020, 12, 2141–2160. [Google Scholar] [CrossRef]
  150. Etemadi, A.; Hemmati, S.; Shahrivar-Gargari, M.; Abibiglue, Y.T.; Bavili, A.; Hamzeh-Mivehroud, M.; Dastmalchi, S. Design, Synthesis, and Biological Evaluation of Novel Indanone Derivatives as Cholinesterase Inhibitors for Potential Use in Alzheimer’s Disease. Chem. Biodivers. 2023, 20, e202300075. [Google Scholar] [CrossRef]
  151. Treuer, A.V.; Faundez, M.; Ebensperger, R.; Hovelmeyer, E.; Vergara-Jaque, A.; Perera-Sardina, Y.; Gutierrez, M.; Fuentealba, R.; Gonzalez, D.R. New NADPH Oxidase 2 Inhibitors Display Potent Activity against Oxidative Stress by Targeting p22(phox)-p47(phox) Interactions. Antioxidants 2023, 12, 1441. [Google Scholar] [CrossRef]
  152. Develin, A.M.; Fuglestad, B. Inositol Hexaphosphate as an Inhibitor and Potential Regulator of p47(phox) Membrane Anchoring. Biochemistry 2024, 63, 1097–1106. [Google Scholar] [CrossRef] [PubMed]
  153. Fan, L.M.; Liu, F.; Du, J.; Geng, L.; Li, J.M. Inhibition of endothelial Nox2 activation by LMH001 protects mice from angiotensin II-induced vascular oxidative stress, hypertension and aortic aneurysm. Redox Biol. 2022, 51, 102269. [Google Scholar] [CrossRef] [PubMed]
  154. Zang, J.; Cambet, Y.; Jaquet, V.; Bach, A. Chemical synthesis of a reported p47phox/p22phox inhibitor and characterization of its instability and irreproducible activity. Front. Pharmacol. 2022, 13, 1075328. [Google Scholar] [CrossRef] [PubMed]
  155. Zang, J.; Peters, F.; Cambet, Y.; Cifuentes-Pagano, E.; Hissabu, M.M.S.; Dustin, C.M.; Svensson, L.H.; Olesen, M.M.; Poulsen, M.F.L.; Jacobsen, S.; et al. Targeting NOX2 with Bivalent Small-Molecule p47phox-p22phox Inhibitors. J. Med. Chem. 2023, 66, 14963–15005. [Google Scholar] [CrossRef]
  156. Trevelin, S.C.; Shah, A.M.; Lombardi, G. Beyond bacterial killing: NADPH oxidase 2 is an immunomodulator. Immunol. Lett. 2020, 221, 39–48. [Google Scholar] [CrossRef]
  157. Glorieux, C.; Liu, S.; Trachootham, D.; Huang, P. Targeting ROS in cancer: Rationale and strategies. Nat. Rev. Drug Discov. 2024, 23, 583–606. [Google Scholar] [CrossRef]
  158. Poolman, T.M.; Quinn, P.A.; Ng, L. Protein Kinase C Epsilon Contributes to NADPH Oxidase Activation in a Pre-Eclampsia Lymphoblast Cell Model. Diseases 2013, 1, 1–17. [Google Scholar] [CrossRef]
  159. Chang, S.N.; Dey, D.K.; Oh, S.T.; Kong, W.H.; Cho, K.H.; Al-Olayan, E.M.; Hwang, B.S.; Kang, S.C.; Park, J.G. Phorbol 12-Myristate 13-Acetate Induced Toxicity Study and the Role of Tangeretin in Abrogating HIF-1alpha-NF-kappaB Crosstalk In Vitro and In Vivo. Int. J. Mol. Sci. 2020, 21, 9261. [Google Scholar] [CrossRef]
  160. Vasukutty, A.; Pillarisetti, S.; Choi, J.; Kang, S.H.; Park, I.K. CXCR4 Targeting Nanoplatform for Transcriptional Activation of Latent HIV-1 Infected T Cells. ACS Appl. Bio Mater. 2024, 7, 4831–4842. [Google Scholar] [CrossRef]
  161. Makni-Maalej, K.; Boussetta, T.; Hurtado-Nedelec, M.; Belambri, S.A.; Gougerot-Pocidalo, M.A.; El-Benna, J. The TLR7/8 agonist CL097 primes N-formyl-methionyl-leucyl-phenylalanine-stimulated NADPH oxidase activation in human neutrophils: Critical role of p47phox phosphorylation and the proline isomerase Pin1. J. Immunol. 2012, 189, 4657–4665. [Google Scholar] [CrossRef]
  162. Kubli-Garfias, C.; Vazquez-Ramirez, R.; Trejo-Munoz, C.; Berber, A. Insights on the mechanism of action of immunostimulants in relation to their pharmacological potency. The effects of imidazoquinolines on TLR8. PLoS ONE 2017, 12, e0178846. [Google Scholar] [CrossRef]
  163. Cheng, N.; He, R.; Tian, J.; Dinauer, M.C.; Ye, R.D. A critical role of protein kinase C delta activation loop phosphorylation in formyl-methionyl-leucyl-phenylalanine-induced phosphorylation of p47(phox) and rapid activation of nicotinamide adenine dinucleotide phosphate oxidase. J. Immunol. 2007, 179, 7720–7728. [Google Scholar] [CrossRef]
  164. Schejtman, A.; Aragao-Filho, W.C.; Clare, S.; Zinicola, M.; Weisser, M.; Burns, S.O.; Booth, C.; Gaspar, H.B.; Thomas, D.C.; Condino-Neto, A.; et al. Lentiviral gene therapy rescues p47(phox) chronic granulomatous disease and the ability to fight Salmonella infection in mice. Gene Ther. 2020, 27, 459–469. [Google Scholar] [CrossRef] [PubMed]
  165. Diebold, B.A.; Fowler, B.; Lu, J.; Dinauer, M.C.; Bokoch, G.M. Antagonistic cross-talk between Rac and Cdc42 GTPases regulates generation of reactive oxygen species. J. Biol. Chem. 2004, 279, 28136–28142. [Google Scholar] [CrossRef] [PubMed]
  166. Vermot, A.; Petit-Hartlein, I.; Smith, S.M.E.; Fieschi, F. NADPH Oxidases (NOX): An Overview from Discovery, Molecular Mechanisms to Physiology and Pathology. Antioxidants 2021, 10, 890. [Google Scholar] [CrossRef] [PubMed]
  167. Kuihon, S.; Sevart, B.J.; Abbey, C.A.; Bayless, K.J.; Chen, B. The NADPH oxidase 2 subunit p47(phox) binds to the WAVE regulatory complex and p22(phox) in a mutually exclusive manner. J. Biol. Chem. 2024, 300, 107130. [Google Scholar] [CrossRef]
  168. Liu, L.; Chen, A.; Li, Y.; Mulder, J.; Heyn, H.; Xu, X. Spatiotemporal omics for biology and medicine. Cell 2024, 187, 4488–4519. [Google Scholar] [CrossRef]
  169. Sanati, M.; Amin Yavari, S. Liposome-integrated hydrogel hybrids: Promising platforms for cancer therapy and tissue regeneration. J. Control. Release 2024, 368, 703–727. [Google Scholar] [CrossRef]
  170. Yu, Y.; Wu, X.; Wang, M.; Liu, W.; Zhang, L.; Zhang, Y.; Hu, Z.; Zhou, X.; Jiang, W.; Zou, Q.; et al. Optogenetic-controlled immunotherapeutic designer cells for post-surgical cancer immunotherapy. Nat. Commun. 2022, 13, 6357. [Google Scholar] [CrossRef]
  171. Klatt, D.; Cheng, E.; Philipp, F.; Selich, A.; Dahlke, J.; Schmidt, R.E.; Schott, J.W.; Buning, H.; Hoffmann, D.; Thrasher, A.J.; et al. Targeted Repair of p47-CGD in iPSCs by CRISPR/Cas9: Functional Correction without Cleavage in the Highly Homologous Pseudogenes. Stem Cell Rep. 2019, 13, 590–598. [Google Scholar] [CrossRef] [PubMed]
  172. Xiong, B.; Zhang, Y.; Liu, S.; Liao, S.; Zhou, Z.; He, Q.; Zhou, Y. NOX Family: Regulators of Reactive Oxygen Species Balance in Tumor Cells. FASEB J. 2025, 39, e70565. [Google Scholar] [CrossRef] [PubMed]
  173. Chen, F.; Li, X.; Aquadro, E.; Haigh, S.; Zhou, J.; Stepp, D.W.; Weintraub, N.L.; Barman, S.A.; Fulton, D.J.R. Inhibition of histone deacetylase reduces transcription of NADPH oxidases and ROS production and ameliorates pulmonary arterial hypertension. Free Radic. Biol. Med. 2016, 99, 167–178. [Google Scholar] [CrossRef] [PubMed]
  174. Deneault, E. Recent Therapeutic Gene Editing Applications to Genetic Disorders. Curr. Issues Mol. Biol. 2024, 46, 4147–4185. [Google Scholar] [CrossRef]
Figure 1. Structure of human p47phox. p47phox contains one phox homology (PX) domain, two SRC homology (SH3) domains, one autoinhibitory region (AIR), and a proline-rich region (PRR). Several serines (marked in red) in the C-terminal region may be phosphorylated to induce activation.
Figure 1. Structure of human p47phox. p47phox contains one phox homology (PX) domain, two SRC homology (SH3) domains, one autoinhibitory region (AIR), and a proline-rich region (PRR). Several serines (marked in red) in the C-terminal region may be phosphorylated to induce activation.
Cells 14 01043 g001
Figure 2. Schematic representation of NOX2. p47phox, p67phox and p40phox remain in the cytosol in a resting state (left side). Upon cell stimulation, p47phox, p67phox and p40phox along with the small G-protein Rac attach to membrane-bound cytochrome b558 (gp91phox and p22phox), and NOX2 becomes activated. Phosphorylation of p47phox is a key step in the activation process that induces the conformational changes required for activation and production of superoxide.
Figure 2. Schematic representation of NOX2. p47phox, p67phox and p40phox remain in the cytosol in a resting state (left side). Upon cell stimulation, p47phox, p67phox and p40phox along with the small G-protein Rac attach to membrane-bound cytochrome b558 (gp91phox and p22phox), and NOX2 becomes activated. Phosphorylation of p47phox is a key step in the activation process that induces the conformational changes required for activation and production of superoxide.
Cells 14 01043 g002
Figure 3. Chemical structures of representative NOX2 pathway modulators.
Figure 3. Chemical structures of representative NOX2 pathway modulators.
Cells 14 01043 g003
Table 1. Regulation of p47phox functional activity by kinases and post-translational modifications.
Table 1. Regulation of p47phox functional activity by kinases and post-translational modifications.
RegulatorSpecific Target(s)Effect on NOX2 ActivationRelevance to Disease States
Protein Kinase C
(PKC)
Ser303, Ser304,
Ser315, Ser320,
Ser328
Phosphorylation at these sites
induces conformational changes
in p47phox, facilitating its
interaction with p22phox and
activation of NOX2
Implicated in diabetic
complications and cancer
progression due to enhanced ROS production
Akt
(Protein Kinase B)
Ser304, Ser328Phosphorylation by Akt sustains
NOX2 activation, prolonging
ROS output
Associated with
neurodegenerative diseases
where prolonged ROS
contributes to neuronal damage
p38 MAPKSer345Phosphorylation at Ser345 primes p47phox for subsequent
phosphorylation events,
enhancing NOX2 activation
Relevant in inflammatory
disorders such as rheumatoid
arthritis and inflammatory
bowel disease
UbiquitinationLysine residues
(specific sites not fully
characterized)
Ubiquitination targets p47phox
for proteasomal degradation,
reducing its availability for
NOX2 complex assembly
Modulates severity in CGD
by regulating p47phox levels
AcetylationLysine residues
(specific sites not fully
characterized)
Acetylation may influence
p47phox stability and
interaction with other proteins,
potentially affecting NOX2
activity
Emerging role in metabolic
stress responses; further
research needed to elucidate
specific mechanisms
Table 2. p47phox’s role in disease pathogenesis.
Table 2. p47phox’s role in disease pathogenesis.
Diseasep47phox RoleImpact of NCF1 DeficiencyPreclinical Outcomes (NCF1/ Models)Human Evidence
Diabetic NephropathyIncreases glomerular ROSReduction in ROS↓ Albuminuria,
improved GFR
↑ NCF1 in renal biopsies
↑ 4-HNE
Diabetic RetinopathyDrives vascular ROSReduction in ROS↓ Vascular leakage,
capillary occlusion
↑ NCF1 in retinal endothelium, ↑ nitrotyrosine
Diabetic NeuropathyPromotes nerve ROSReduction in ROS↑ Nerve conduction velocity, ↓ pain
hypersensitivity
↑ NCF1 in peripheral nerves, ↑ lipid peroxidation
NAFLD/ObesityAmplifies hepatic/adipose ROSReduction in hepatic ROS↓ Lipid accumulation, ↓ fibrosis↑ NCF1 in visceral adipose
↑ ROS markers
Chronic
Granulomatous
Disease
(CGD)
Impairs neutrophil ROS productionReduction in ROS↑ Susceptibility to
Aspergillus and
S. aureus, ↓ microbial clearance
↓ NCF1 mutations in PBMCs, ↑ infection rates, ↓ superoxide levels
Rheumatoid
Arthritis
(RA)
Drives synovial ROSReduction in ROS↓ Synovial inflammation,
↓ cartilage erosion
↑ NCF1 in synovial biopsies, ↑ malondialdehyde, ↑ IL-6
Inflammatory
Bowel
Disease
(Crohn’s Disease)
Supports gut microbial ROSReduction in ROS↑ Gut permeability, ↑ Clostridium overgrowth, ↑ colitis severity↓ NCF1 in Crohn’s flares, ↑ IL-1β, ↓ microbial control
Inflammatory
Bowel
Disease
(Ulcerative Colitis)
Amplifies mucosal ROSReduction in ROS↓ Mucosal damage, ↓
inflammatory cytokines
↑ NCF1 in severe UC, ↑ ROS markers, ↑ TNF-α
Systemic
Lupus
Erythematosus
(SLE)
Promotes autoantigen ROSReduction in ROS↓ Immune complex deposition, ↓ renal damage↑ NCF1 in lupus nephritis, ↑ 8-OHdG, ↑ anti-dsDNA antibodies
MelanomaEnhances immunosuppressive ROSReduction in ROS↓ Tumor growth, ↑ T-cell infiltration by ↑ NCF1 in tumor-associated macrophages, ↑ MDSC markers
LymphomaDrives tumor-promoting ROSReduction in ROS↑ Tumor cell lysis, ↓ tumor burden↑ NCF1 in lymphoma biopsies, ↑ ROS-driven immunosuppression
Pancreatic CancerPromotes tumoricidal resistance ROSReduction in ROS↑ Macrophage tumoricidal activity, ↓ tumor burden↑ NCF1 in pancreatic tumors, ↑ ROS-driven PD-L1 expression
Breast CancerEnhances tumor-promoting ROSReduction in ROS↓ Tumor progression, ↑ immune cell infiltration ↑ NCF1 in breast tumor stroma, ↑ ROS-driven VEGF expression
Colorectal CancerDrives immunosuppressive ROSReduction in ROS↓ Tumor growth, ↓ metastasis ↑ NCF1 in colorectal tumors, ↑ ROS-driven MDSC activity
PsoriasisPromotes keratinocyte ROSReduction in ROS↓ Epidermal hyperplasia; ↓ inflammatory infiltrationNCF1 mutations associated with increased susceptibility
AsthmaEnhances airway epithelial ROSReduction in ROSAirway inflammation; ↓ eosinophil infiltration↑ NCF1 expression in bronchial biopsies; ↑ oxidative stress markers
Multiple Sclerosis (MS)Mediates CNS inflammation via ROSReduction in ROS↓ Demyelination; ↓ neuroinflammation↑ NCF1 expression in active lesions; ↑ oxidative damage markers
Parkinson’s DiseaseContributes to dopaminergic neuron degeneration via ROSReduction in ROS↓ Neuronal loss; ↑ motor function↑ NCF1 in substantia nigra; ↑ oxidative stress markers
Table 3. Disease-specific signaling pathways involving p47phox.
Table 3. Disease-specific signaling pathways involving p47phox.
DiseaseSignaling
Pathway(s)
ROS-Mediated EffectDownstream
Consequences
Chronic
Granulomatous
Disease
(CGD)
Impaired NF-κB, IL-1βDeficient ROS production leading to impaired microbial killingRecurrent bacterial and fungal infections, granuloma formation, hyperinflammation
Cardiovascular
Diseases
TGF-β, NF-κBOxidation of LDL, promotion of atherosclerotic plaque formationEndothelial dysfunction, vascular inflammation, plaque progression
Neurodegenerative DiseasesPI3K/Akt, MAPKEnhanced microglial activation, increased α-synuclein aggregationDopaminergic neuronal loss, cognitive decline
Inflammatory
Disorders
NF-κB, IL-6, TNF-αAmplification of synovial inflammation via ROS productionJoint destruction, cartilage degradation
Metabolic
Disorders
PKC, MAPK, TGF-βInduction of glomerular oxidative stress and fibrosisProgression of diabetic nephropathy and retinopathy
CancerPI3K/Akt, Wnt/β-cateninPromotion of epithelial-mesenchymal transition (EMT), increased proliferationTumor growth, invasion, metastasis
Table 4. p47phox-targeted therapeutics.
Table 4. p47phox-targeted therapeutics.
StrategyAgent/ExampleTarget/MechanismPreclinical EfficacyChallenges
Inhibitors
(Peptides)
NOX2ds-tatInhibit p47phox-p22phox PPIReduced vascular ROS and inflammation; improved hind-limb perfusion in ischemic models.Rapid clearance, limited bioavailability
Inhibitors
(Small Molecules)
CPP11G
CPP11H
Celastrol
Ebselen/SPI-1005
LMH001
C6
C14
Bivalent inhibitors
Disrupt p47phox-PPI
Redox modulation
Decreased synovial ROS; neuroprotection in models of neurodegenerationOff-target effects; hepatotoxicity concerns
Agonists
(Gene Therapy)
Lentiviral NCF1
PM359
Restore NCF1 expression to reconstitute NADPH oxidase activityReduced infection rates and granuloma formation in CGD mouse modelsPotential genomic instability; variable transduction efficiency
Agonists
(PKC Activators)
Phorbol 12-myristate 13-acetate
(PMA)
Induce p47phox phosphorylation to activate NADPH oxidaseEnhanced microbial clearance; increased tumor cell lysis in vitroSystemic toxicity; pro-inflammatory responses
Agonists
(TLR Ligands)
CL097Activate TLR7/8 pathways; modulate Pin1 and Ser345 phosphorylation of p47phoxStimulated plasmacytoid dendritic cell activation; increased cytokine production; potential antitumor effectsShort half-life; challenges in targeted delivery
Table 5. Clinical trial status of NOX-targeted therapies.
Table 5. Clinical trial status of NOX-targeted therapies.
Clinical Development
Stage
Agent/TargetDisease IndicationClinicaltrials.gov
NCT #
Phase IISentanaxib
(GKT137831)
NOX1/4 inhibitor
Idiopathic pulmonary fibrosis
Type 2 diabetes & albuminuria
Biliary cholangitis
Alport syndrome
NCT03865927
NCT02010242
NCT03226067
NCT06274489
Phase IIAPX-115
Pan-NOX inhibitor
Type 2 diabetes with nephropathy
COVID-19
Acute kidney injury
NCT04534439
NCT04880109
NCT05758896
Phase II/IIISPI-1005/Ebselen
NOX2 inhibitor
Bipolar disorder
Treatment-resistant depression
Meniere’s disease
Cochlear implant
Type I & II diabetes
COVID-19
Otoprotection
NCT03013400
NCT05117710
NCT06859788, NCT03325790
NCT06340633
NCT00762671
NCT04484025, NCT04483973
NCT01444846, NCT01451853, NCT02779192
Phase IPM359
NCF1 gene therapy
Chronic granulomatous diseaseNCT06559176
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gamble, M.E.; Sureshkumar, S.; Carrera Espinoza, M.J.; Hakim, N.L.; Espitia, C.M.; Bi, F.; Kelly, K.R.; Wang, W.; Nawrocki, S.T.; Carew, J.S. p47phox: A Central Regulator of NADPH Oxidase Function and a Promising Therapeutic Target in Redox-Related Diseases. Cells 2025, 14, 1043. https://doi.org/10.3390/cells14141043

AMA Style

Gamble ME, Sureshkumar S, Carrera Espinoza MJ, Hakim NL, Espitia CM, Bi F, Kelly KR, Wang W, Nawrocki ST, Carew JS. p47phox: A Central Regulator of NADPH Oxidase Function and a Promising Therapeutic Target in Redox-Related Diseases. Cells. 2025; 14(14):1043. https://doi.org/10.3390/cells14141043

Chicago/Turabian Style

Gamble, Madison E., Sruthi Sureshkumar, Maria Janina Carrera Espinoza, Natalie L. Hakim, Claudia M. Espitia, Fangchao Bi, Kevin R. Kelly, Wei Wang, Steffan T. Nawrocki, and Jennifer S. Carew. 2025. "p47phox: A Central Regulator of NADPH Oxidase Function and a Promising Therapeutic Target in Redox-Related Diseases" Cells 14, no. 14: 1043. https://doi.org/10.3390/cells14141043

APA Style

Gamble, M. E., Sureshkumar, S., Carrera Espinoza, M. J., Hakim, N. L., Espitia, C. M., Bi, F., Kelly, K. R., Wang, W., Nawrocki, S. T., & Carew, J. S. (2025). p47phox: A Central Regulator of NADPH Oxidase Function and a Promising Therapeutic Target in Redox-Related Diseases. Cells, 14(14), 1043. https://doi.org/10.3390/cells14141043

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop