Next Article in Journal
Propolis Extract with Activity Against Cutibacterium acnes Biofilm Targeting the Expression of Virulence Genes
Previous Article in Journal
Hydnocarpin, a Natural Flavonolignan, Induces the ROS-Mediated Apoptosis of Ovarian Cancer Cells and Reprograms Tumor-Associated Immune Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Manganese Superoxide Dismutase: Structure, Function, and Implications in Human Disease

by
Jovan Grujicic
1,2 and
Antiño R. Allen
1,2,3,4,*
1
Division of Radiation Health, University of Arkansas for Medical Sciences, Little Rock, AR 72205, USA
2
Department of Pharmacology and Toxicology, University of Arkansas for Medical Sciences, Little Rock, AR 72205, USA
3
Department of Pharmaceutical Sciences, University of Arkansas for Medical Sciences, Little Rock, AR 72205, USA
4
Neurobiology & Developmental Sciences, University of Arkansas for Medical Sciences, Little Rock, AR 72205, USA
*
Author to whom correspondence should be addressed.
Antioxidants 2025, 14(7), 848; https://doi.org/10.3390/antiox14070848
Submission received: 24 June 2025 / Revised: 7 July 2025 / Accepted: 8 July 2025 / Published: 10 July 2025
(This article belongs to the Section Antioxidant Enzyme Systems)

Abstract

Manganese superoxide dismutase (MnSOD) is a vital mitochondrial antioxidant enzyme that preserves cellular integrity by catalyzing the dismutation of superoxide radicals into hydrogen peroxide. Its central role in maintaining redox homeostasis has positioned it as a key target in biomedical research. This review provides an in-depth examination of MnSOD’s structural and functional properties, regulatory mechanisms, and its involvement in the pathogenesis of various human diseases.

1. Introduction

Mitochondria are multifunctional organelles essential to processes such as ATP generation, calcium homeostasis, metabolite biosynthesis, and apoptosis regulation [1]. Among these, ATP production via oxidative phosphorylation is perhaps the most vital for life; however, it also results in reactive oxygen species (ROS) formation, primarily as superoxide anions (O2•−) due to electron leakage from the electron transport chain [2]. While physiological ROS levels play signaling roles, their excessive accumulation disrupts the redox balance and damages biomolecules, contributing to various diseases [1].
Superoxide dismutases (SODs) are metalloenzymes that catalyze the dismutation of O2•− into oxygen and hydrogen peroxide (H2O2), preventing oxidative damage [3]. Based on their metal cofactors, SODs are classified into copper zinc SOD (SOD1), iron SOD (SOD3), manganese (Mn)SOD, and nickel SOD isoforms [4]. These enzymes are compartmentalized for optimal function: SOD1 is cytosolic and enriched in tissues like liver and kidney; SOD3 is extracellular in lower organisms but also found in chloroplasts [5,6]. In humans, SOD3 is largely confined to extracellular sites with limited expression.
MnSOD, localized in the mitochondrial matrix, is the primary antioxidant defense within mitochondria. When redox homeostasis is disturbed (e.g., genetic mutations, environmental toxins, or therapeutic side effects), oxidative stress develops. This stress contributes to the onset and progression of diseases such as cancer, cardiovascular disease, renal and neurodegenerative disorders, and obesity. These conditions can further enhance ROS production, forming a positive feedback loop that amplifies oxidative damage (Figure 1). Therefore, MnSOD is especially important in energy-demanding organs such as the heart, brain, and liver, where mitochondrial ROS production is highest [7].
Due to their essential role in oxidative homeostasis, SODs have emerged as key targets in research on inflammation, cancer, aging, and metabolic dysfunction. MnSOD has garnered attention for its pivotal role in preserving mitochondrial integrity under oxidative stress. This review aims to provide an updated overview of MnSOD’s structure, function, and relevance to human disease, offering insight into its therapeutic potential.

2. MnSOD

MnSOD is exclusively located within the mitochondrial matrix and is the only SOD within the mitochondria [8,9]. MnSOD defends against oxidative stress by converting O2•− into H2O2 and oxygen. In 1969, the enzymatic activity of SODs was discovered when a copper protein was identified that could catalytically remove the Pauling free radical, the first experimentally recognized form of superoxide [10,11]. This discovery marked the first clear evidence that superoxide was a biologically relevant free radical and led to the identification of SODs as essential antioxidant enzymes. MnSOD was first isolated in E. coli in 1970 and was initially thought to be a homodimer of approximately 40 kDa, with each subunit weighing approximately 20 kDa [12]. In 1972, mitochondrial MnSOD was isolated from chicken liver with a calculated molecular weight of 79,400 (twice that of bacterial SOD), comprising four identical subunits, each weighing around 20,200 kDa, making it a homotetramer. The human MnSOD, encoded by MnSOD on chromosome 6q25.3, has a highly conserved sequence with over 40% homology shared among the human, yeast, and E. coli proteins [13,14]. The function of this homotetramer is essential for normal cellular metabolism, and its deficiency or dysfunction is linked to various diseases, including neurodegenerative, kidney, and cardiovascular conditions [7,15,16,17]. MnSOD is regulated through transcriptional, epigenetic, and post-translational modifications to ensure appropriate activity in response to oxidative stress and environmental factors [17,18].

2.1. Structure and Mechanism of Action

MnSOD operates as a tetrameric enzyme, composed of four identical subunits, each housing a catalytically essential manganese ion (Mn2+/Mn3+) within its active site. This tetrameric organization enhances structural stability and promotes intersubunit interactions that are key to the enzymatic efficiency of MnSOD [19,20]. This structural feature is exemplified in S. cerevisiae MnSOD, which shows heightened resistance to thermal and chemical unfolding compared to the dimeric C. albicans form, which more readily dissociates when crucial interfacial residues are altered [21]. In human MnSOD, residues glutamic acid (Glu)162 and histidine (His)30 are crucial at the dimer interface, where they form hydrogen bonds that help maintain structural integrity and catalytic competence [22,23]. Substitution of Glu162 has been shown to reduce MnSOD’s enzymatic activity by as much as 95% [24].
At the catalytic core, the manganese ion is ligated by inner sphere residues His26, His74, His163, and aspartic acid (Asp)159, along with a coordinating solvent molecule (referred to as WAT1, the first-shell water ligand bound to the manganese ion) that may exist as water or hydroxide [23]. The coordination geometry of the Mn2+ and Mn3+ active sites is illustrated in Figure 2. This configuration is essential for SOD [25]. Supporting this core, outer sphere residues including His30, tyrosine (Tyr)34, phenylalanine (Phe)77, tryptophan (Trp)78, Trp123, glutamine (Gln)143, Trp161, and Glu162 form a hydrogen bond network that facilitates proton transfer and stabilizes the catalytic transition states [22,23].
McAdam et al. described a four-step catalytic cycle for MnSOD consisting of a fast outer sphere pathway and a slower inner sphere route (Table 1) [26]. In both pathways, the cycle begins with the reduction of Mn3+ to Mn2+ by the first O2•−, accompanied by protonation of the coordinated hydroxide by Gln143 to form water (Step 1, rate constant k1) [26,27]. In the fast pathway, a second O2•− oxidizes Mn2+ back to Mn3+ while generating H2O2 via interaction with Tyr34 and WAT1 (Step 2, k2) [26,28]. Under conditions of high O2•− or low temperature, a reversible peroxide adduct forms with Mn2+ (Step 3, k3), temporarily inhibiting enzyme activity. The resolution of this product-inhibited complex requires two protons, delivered by WAT1 and surrounding outer sphere residues, to regenerate Mn3+ and release H2O2 (Step 4, k4) [20]. The ratio of k2/k3, referred to as the gating ratio, determines which pathway predominates [26].
Azadmanesh et al. later proposed a detailed mechanism centered on concerted proton-electron transfer (CPET), in which WAT1 and residues such as His26, His74, His163, Asp159, and Gln143 mediate coupled redox transitions of the manganese ion [25]. In the Mn3+ state, WAT1 is deprotonated and hydrogen-bonded to Gln143 and Asp159. Upon reduction to Mn2+, WAT1 is reprotonated via Gln143, which stabilizes the reduced state by forming strong hydrogen bonds with Trp123 and WAT1. A sixth ligand, possibly a hydroxide ion, also participates in stabilizing the Mn2+ conformation and facilitating the next catalytic round.
Recently, direct ultrafast crystallographic snapshots provided experimental confirmation of this CPET model. The study captured, at microsecond resolution, the real-time proton transfer from Gln143 to WAT1 and conformational shifts involving Trp123 as Mn transitioned between oxidation states. These findings validate the dynamic proton relay system involving Gln143–WAT1–Trp123 that drives MnSOD catalysis [29].
Temperature plays a crucial role in modulating these steps by influencing the gating ratio. While the initial steps (k1, k2) are relatively temperature-insensitive, the rates of peroxide adduct formation and its resolution (k3, k4) increase with temperature, peaking around 55 °C [26]. This shift favors the slower, inhibited cycle. Conversely, lower temperatures reduce enzyme inhibition, maintaining rapid turnover via the fast pathway [30].

2.2. Transcriptional and Post-Translational Regulation of MnSOD

Transcriptional regulation and post-translational modifications can modulate the activity and stability of the enzyme. Three main regions of SOD2 are involved in regulating MnSOD expression: the 5′ upstream enhancer, the proximal promoter, and the intronic enhancer [31]. The 5′ upstream enhancer region contains binding sites for many transcription factors, namely activator protein-1 (AP-1), cAMP response element-binding protein (CREB), nuclear factor erythroid 2-related factor 2 (Nrf2), forkhead box O3a (FoxO3a), nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB), and hypoxia-inducible factor 1-alpha (HIF-1α) [16]. Nrf2 binds to the antioxidant response element in the promoter region of antioxidant genes, consequently increasing MnSOD expression; similarly, PGC-1α promotes the expression of antioxidative enzymes, including MnSOD [16]. Additionally, PGC-1α induces Nrf2-mediated MnSOD expression in the liver during Staphylococcus aureus-induced peritonitis [32]. HIF-1α and protein kinase B (Akt)-induced phosphorylation of forkhead box O3a (FoxO3a) downregulates MnSOD transcription, whereas sirtuin 3 (SIRT3), a mitochondrial nicotinamide adenine dinucleotide (NAD+)-dependent deacetylase, deacetylates FoxO3a and has the opposite effect [16]. The proximal promoter of MnSOD has CG repeats with binding sites for specificity protein 1 (Sp1), a transcriptional activator that promotes MnSOD expression. Activating protein-2 (AP-2), a transcription factor, also binds to the proximal promoter and competes with Sp1 for binding to the same promoter regions. When AP-2 binds, it can either directly inhibit Sp1 by occupying these sites or form AP-2/Sp1 complexes. The formation of AP-2/Sp1 complexes sequesters Sp1 away from the promoter, preventing it from activating MnSOD transcription effectively, thus reducing MnSOD expression [31]. Lastly, the intronic enhancer, located within the second intron, enhances Sp1 activity in the proximal promoter and possesses a nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) binding site that interacts with CCAAT/enhancer-binding protein beta (C/EBP-β) to promote MnSOD transcription. Additionally, MnSOD is susceptible to inactivation by peroxynitrite (ONOO), which nitrates critical tyrosine residues [16,31,33,34]. These regulatory mechanisms are summarized in Figure 3.
Post-translational modifications of MnSOD include acetylation, nitration, phosphorylation, and glutathionylation. The acetylation of human MnSOD at lysine residues is regulated by SIRT3 [35,36]. Acetylation reduces SOD activity, resulting in higher mitochondrial O2•− levels, whereas irradiation, nutrient scarcity, and oxidative stress trigger SIRT3-mediated deacetylation of MnSOD to adequately enhance ROS scavenging within mitochondria to combat the emerging challenging conditions [16]. Nitric oxide is an intercellular messenger that has the potential to form ONOO, which can inactivate MnSOD by nitrating Tyr34 at the active site [37]. Phosphorylation and dephosphorylation can either increase or decrease MnSOD activity depending on the specific amino acid residue modified, demonstrating a target site-based effect [16]. Like many other enzymes, permanent oxidation has the ability to pathologically shut down the activity of MnSOD, and S-glutathione, due to its ability to protect thiols from irreversible oxidation, can modulate MnSOD activity [16]. These post-translational modifications are also included in the schematic overview presented in Figure 3.

2.3. MnSOD as a Thermoreceptor

A recently proposed function of MnSOD is as a catalytic thermoreceptor. At lower temperatures, the enzyme increasingly favors the fast catalytic cycle, and when the temperature drops sufficiently, all reactions shift exclusively to this pathway, resulting in a marked increase in enzymatic activity [30]. This shift defines a threshold below which MnSOD becomes specifically activated by cold. For example, at 25 °C, the gating ratio of human MnSOD is approximately 1, while at 20 °C, it rises to 4, indicating a strong preference for the fast cycle under cold conditions [38].
Cold activation of MnSOD represents a proactive cellular adaptation to low temperatures, as the enzyme’s activity is upregulated beyond what is required to manage baseline superoxide levels. Kinetic modeling indicates that lower mitochondrial superoxide concentrations can drive the coenzyme Q reaction forward, increasing electron leakage and thereby amplifying H2O2 production by MnSOD [30,39]. MnSOD plays a central role in balancing superoxide and H2O2 fluxes to preserve redox homeostasis, a function that becomes especially crucial during temperature fluctuations. Under cold conditions, MnSOD not only accelerates SOD reactions via the fast cycle but also contributes to temperature regulation by promoting thermogenesis and ATP production, enabling efficient heat and energy generation. In contrast, higher temperatures favor the slower, product-inhibited cycle, potentially resulting in superoxide accumulation. By shifting toward the fast cycle at low temperatures, MnSOD enhances ROS detoxification and the thermogenic conversion of nicotinamide adenine dinucleotide (NADH) and flavin adenine dinucleotide (FADH2), highlighting its essential role in maintaining cellular energy and temperature balance during adaptive thermogenesis [30]. Excess H2O2 from cold-activated MnSOD may signal low temperatures and trigger physiological responses [30]. H2O2, a product of MnSOD activity, acts as a signaling molecule in energy metabolism and thermogenesis, modulating the uncoupling proteins by desensitizing them to inhibition and activating them through modifications like sulfenylation of UCP1-Cys253 and glutathionylation of UCP2-3 [40]. Additionally, H2O2 activates mitochondrial phospholipase A2γ, releasing free fatty acids that promote superoxide-induced uncoupling and increase cellular respiration; glutathione counters these effects [41,42].
The interaction between H2O2 and glutathione highlights how redox balance plays an important role in thermogenesis. In mammals, brown adipose tissue produces heat using the uncoupling proteins that are regulated by temperature and norepinephrine signals [40]. Some evidence suggests that transient receptor potential channels help sense temperature and may be influenced by ROS, although the details are still unclear [43,44]. When MnSOD is deleted, superoxide levels increase while H2O2 decreases, which impairs the body’s ability to generate heat. This finding supports the idea that both MnSOD and H2O2 are important for proper thermogenic responses [30]. Because MnSOD is found in all cells and responds to temperature shifts, it has been suggested that it might act as a general thermosensor across different tissues [30].

2.4. MnSOD and Obesity

MnSOD is essential for regulating mitochondrial function and energy metabolism in adipocytes, and its dysfunction has far-reaching consequences for obesity and metabolic health. Alterations in MnSOD expression or activity, whether genetic, diet-induced, or tissue-specific, can influence adiposity, insulin sensitivity, and oxidative stress. In adipocyte-specific SOD2 knockout mice that were fed a high-fat diet, MnSOD deletion surprisingly led to reduced fat mass, smaller adipocytes, lower leptin levels, and improved glucose and insulin tolerance, despite elevated superoxide levels and no major upregulation of compensatory antioxidant enzymes [45]. These mice also showed enhanced fatty acid oxidation, increased mitochondrial biogenesis, and protection from hepatic steatosis and inflammation, suggesting that adipocyte MnSOD normally supports energy storage and that its absence may favor energy expenditure through mitochondrial uncoupling.
Conversely, systemic MnSOD impairment appears to promote obesity. A study of 815 elderly individuals found that the Val/Val genotype of the Ala16Val MnSOD polymorphism—known to impair mitochondrial import of MnSOD—was more common among obese participants, highlighting how genetic disruption of mitochondrial antioxidant defenses may predispose individuals to oxidative stress and fat accumulation, thereby supporting a link between oxidative imbalance and increased adiposity [46]. Similarly, mice lacking MnSOD in intestinal epithelial cells developed spontaneous obesity marked by elevated inflammation, increased lipogenesis, and insulin resistance. These effects were driven by phospholipase A2 activation and excessive arachidonic acid release, but were reversed by an essential fatty acid-deficient diet [47]. A negative correlation between intestinal MnSOD and obesity markers in human samples further supported its translational relevance.
In another setting, skeletal muscle biopsies from obese pregnant women, with and without gestational diabetes, revealed reduced MnSOD and complex II (succinate dehydrogenase) activity compared to normal-weight controls, despite unchanged MnSOD protein levels [48]. Complex II is a key enzyme in both the tricarboxylic acid (TCA) cycle and the electron transport chain, and its impaired activity reflects mitochondrial dysfunction. The diminished activity was linked to increased acetylation of MnSOD due to reduced SIRT3 expression and activity, implicating impaired deacetylation as a mechanism for reduced antioxidant function. These women also exhibited greater oxidative stress, decreased PGC-1α expression, and a compromised mitochondrial capacity, reinforcing the notion that obesity itself can suppress MnSOD function.
Altogether, these studies reveal a nuanced relationship between MnSOD and obesity: while targeted loss of MnSOD in adipocytes may promote leanness and metabolic benefits, systemic or nonadipose MnSOD deficiency tends to drive inflammation, insulin resistance, and fat accumulation. Moreover, obesity may itself reduce MnSOD function, creating a feedback loop that worsens oxidative stress and mitochondrial dysfunction. These findings highlight the potential of MnSOD and its regulatory pathways as therapeutic targets for obesity-related metabolic disorders, though tissue-specific strategies will be crucial moving forward.

2.5. MnSOD and Renal Disease

The kidney is a metabolically demanding organ with high mitochondrial oxidative activity, particularly within the proximal tubular cells responsible for nutrient reabsorption after glomerular filtration. This makes it especially susceptible to oxidative stress-induced damage [16]. In particular, diabetes—a leading cause of chronic kidney disease—is closely associated with mitochondrial oxidative stress and MnSOD dysfunction. Mitochondrial oxidative stress has been directly implicated in the development and progression of chronic kidney disease and acute kidney injury, as elevated mitochondrial ROS can drive mitochondrial dysfunction, ATP depletion, inflammation, and apoptosis [16,49,50]. MnSOD plays a protective role in maintaining renal health. MnSOD-deficient mice show elevated oxidative stress, tubular injury, glomerular sclerosis, and increased infiltration of renal interstitial T cells and macrophages compared to wild-type controls [51]. In ischemia/reperfusion acute kidney injury models, oxidative stress leads to MnSOD nitration and deactivation, contributing to mitochondrial damage [52,53]. Even when kidney-specific MnSOD knockout mice maintain overall renal function, localized damage occurs, underscoring MnSOD’s importance in sustaining redox balance in renal mitochondria [54]. In albumin-overload models, decreased MnSOD levels are associated with increased mitochondrial oxidative stress and renal inflammation. In chronic renal allograft rejection and diabetic kidney disease, MnSOD dysfunction correlates with disease progression and elevated oxidative stress, and several therapeutic interventions targeting these pathways have shown promise [55,56,57]. Similarly, in membranous nephropathy, Moshen granule treatment alleviated kidney injury by activating the aryl hydrocarbon receptor (AhR), which upregulated Nrf2 and subsequently enhanced MnSOD expression, reducing oxidative stress and inflammation [58]. In a more recent model, proximal tubule-specific deletion of MnSOD in diabetic mice produced stage-dependent effects: early diabetes was associated with improved kidney function via NRF2 activation and enhanced fatty acid oxidation, while late-stage disease showed worsened renal injury due to impaired mitochondrial respiration and heightened inflammation, highlighting MnSOD’s dual role in disease progression [59]. Genetic studies have also linked the Ala16Val polymorphism in the MnSOD gene to diabetic kidney disease [60]. The Val/Val genotype is associated with increased diabetic nephropathy risk in Finnish, Swedish, and Japanese patients, while the Ala/Ala and Ala/Val genotypes are protective in Chinese and Mexican populations [61,62,63,64,65,66,67]. These genotypes differ in how efficiently MnSOD is imported into the mitochondria: the Val variant impairs import by altering the targeting sequence’s secondary structure, reducing mitochondrial MnSOD levels and antioxidant protection, whereas the Ala variant facilitates efficient import and preserves mitochondrial defenses [68].
MnSOD is also deeply involved in diabetes pathology. Type 1 and type 2 diabetes mellitus (T1DM and T2DM, respectively) are worsened by oxidative stress driven by glucose autoxidation [69,70,71]. The CT genotype of the MnSOD 47C/T polymorphism is associated with reduced MnSOD mRNA levels, which in turn increases susceptibility to type 1 diabetes mellitus (T1DM) in newly diagnosed patients [72,73]. In streptozotocin (STZ)-induced T1DM mice, MnSOD overexpression in pancreatic β-cells improved glucose control by inhibiting NF-κB activation and scavenging ROS [74]. Similarly, islets overexpressing MnSOD isolated from non-obese diabetic mice during the prediabetic stage and transplanted into STZ-induced diabetic severe combined immunodeficiency mice resulted in improved survival and prolonged function [75]. Long-term hyperglycemia causes diabetic complications, including nephropathy, neuropathy, and retinopathy, largely through oxidative stress-induced β-cell dysfunction and insulin resistance [70,76]. In diabetic cardiomyopathy, MnSOD overexpression preserves mitochondrial function and contractility by enhancing catalase activity and reducing oxidative damage [77,78]. MnSOD also protects against oxidative stress-induced apoptosis in detrusor smooth muscle during diabetic bladder dysfunction [79] and prevents diabetic retinopathy by reducing glucose-induced mitochondrial DNA damage [80]. Consistent with this, the multifunctional redox-sensitive protein DJ-1 (also known as PARK7) was shown to protect retinal ganglion cells from high glucose-induced oxidative stress by restoring MnSOD expression and activity, thereby maintaining mitochondrial membrane potential and promoting neuronal survival [81].
Recently, the mitochondrial acetyltransferase GCN5L1 was found to be upregulated in diabetic kidney disease, leading to increased MnSOD acetylation and decreased activity, exacerbating oxidative stress and renal injury [82]. Knocking down GCN5L1 restored MnSOD activity, reduced oxidative stress, and improved renal inflammation and fibrosis, suggesting that this pathway could be a therapeutic target [82].
In summary, the kidney’s metabolic intensity makes it vulnerable to mitochondrial oxidative damage; here, MnSOD serves a protective role by preserving redox balance and mitochondrial function [16,51,54]. Its dysfunction is implicated in multiple kidney diseases, including membranous nephropathy and diabetic nephropathy, in which both genetic polymorphisms (such as Ala16Val and 47C/T) and disease stage influence the disease outcomes [58,59,60,72]. MnSOD also protects against a wide range of diabetes-related complications, including cardiomyopathy, retinopathy, and transplant rejection, and is supported by additional regulators such as DJ-1 and GCN5L1, which modulate MnSOD activity under oxidative stress [81,82]. Targeting MnSOD function—whether by enhancing its expression, preventing its inactivation, or regulating upstream effectors—may offer promising therapeutic strategies for both renal and metabolic diseases [82].

2.6. MnSOD in Cardiovascular Disease

The heart is one of the most energy-demanding organs in the body and consequently generates high levels of ROS [83]. This makes it particularly sensitive to impairments in antioxidant defense systems, especially those involving MnSOD. The role of MnSOD in the cardiovascular system is complex, as high and low levels can impact the disease risk differently depending on the tissue and cell type involved. For this reason, MnSOD expression and activity levels have been investigated as predictive biomarkers for cardiovascular conditions, even when those conditions are not directly caused by MnSOD dysfunction [83]. In some cases, elevated MnSOD is protective, while in others, it may reflect an adaptive response to pathology and correlate with worse outcomes [84].
Oxidative stress is a key factor in hypertension, a disease where MnSOD plays a protective role. Oxidative stress contributes to vascular stiffness and smooth muscle cell adhesion, exacerbating endothelial dysfunction and increasing blood pressure [85,86]. Within the brain, MnSOD regulates blood pressure by reducing mitochondrial superoxide anions. Deletion of MnSOD in the subfornical organ raises the mean arterial pressure and intensifies angiotensin II-induced hypertension in mice [87]. Similarly, reduced MnSOD in the brainstem rostral ventrolateral medulla of hypertensive rats leads to elevated ROS and blood pressure, whereas MnSOD gene delivery to this region lowers arterial pressure [88]. Delivering MnSOD to carotid and femoral arteries also mitigates hypertension and delays vascular dysfunction [89].
Findings in neonatal pulmonary hypertension have also implicated MnSOD. Loss of MnSOD function in animal models leads to persistent pulmonary hypertension of the newborn (PPHN), with elevated right ventricular pressure and endothelial apoptosis [90]. Reintroducing MnSOD restores nitric oxide signaling and improves vascular relaxation, indicating therapeutic promise for this approach in PPHN [90,91].
Genetic variations in MnSOD influence the susceptibility to cardiovascular diseases. In a Taiwanese population, the Val/Ala or Ala/Ala MnSOD genotypes were associated with increased coronary artery disease risk and greater severity in patients with one- or two-vessel disease [92]. No such associations were observed for catalase or glutathione peroxidase polymorphisms, making MnSOD an independent risk factor [92]. These results were mirrored in a Tunisian cohort [93].
Cardiomyocyte-specific SOD2 knockout in mice results in early-onset dilated cardiomyopathy and death by four months due to heart failure [92,94]. The mice displayed mitochondrial abnormalities, excessive ROS, impaired respiration, and metabolic shifts without compensatory antioxidant responses [92]. A related inducible model with SOD2+/− and SOD2−/− mice found that global knockout led to early mortality and reduced activity, while heterozygous deletion impaired cardiac function, promoted fibrosis, and heightened sensitivity to doxorubicin-induced oxidative stress [95].
Other cardiovascular structures, such as aortic valves, are also affected by MnSOD levels. Mice with MnSOD haploinsufficiency showed increased oxidative stress and calcification without functional valve changes, whereas MnSOD overexpression enhanced osteogenic markers without lowering the calcium burden, highlighting complex effects depending on the disease stage and cell type [96].
The link between MnSOD and atherosclerosis is well documented. Loss of MnSOD in ApoE-deficient mice accelerates plaque formation, enlarges necrotic cores, and promotes vascular inflammation, especially at arterial branch points prone to high shear stress [97,98,99].
Therapeutically, MnSOD also appears to mediate the protective effects of tamoxifen. In cardiomyocytes, tamoxifen increased MnSOD levels and reduced apoptosis following adriamycin exposure, suggesting that MnSOD induction is one mechanism for its cardioprotective effects [100].
Mitochondrial oxidative stress plays a major role in diabetic cardiomyopathy, where MnSOD overexpression has shown protective effects. In T1DM mice, cardiac-specific SOD2 overexpression preserved mitochondrial structure, restored glutathione levels, and maintained cardiac function, although full respiratory function was not completely restored [78].
Platelets are also affected by MnSOD. In its absence, platelet mitochondria produce excess ROS, which increase platelet activation and aggregation, promoting thrombosis and contributing to cardiovascular disease risk [101]. These effects underscore MnSOD’s broader role in vascular health and inflammation [101]. Supporting this, a reduction in mitochondrial ROS through SOD2-targeted therapy improved oxidative phosphorylation and preserved endothelial resilience in non-reperfused myocardial infarction, highlighting MnSOD’s therapeutic potential in protecting coronary endothelium under ischemic conditions [102].
Atrial fibrillation (AF) may also be linked to MnSOD levels. A study measuring plasma MnSOD in 130 patients found the highest levels in those with paroxysmal AF, with lower levels in persistent AF and the lowest levels in the controls [103]. These levels correlated with atrial size and red blood cell metrics, and logistic regression identified MnSOD as an independent risk factor for paroxysmal AF [103]. However, whether this reflects a compensatory response or causal involvement remains unclear.
When it comes to heart failure prognosis, elevated MnSOD levels have been linked to poor outcomes. In one cohort, the patients who died or required transplant had higher MnSOD activity, which negatively correlated with ejection fraction and renal function while positively correlating with N-terminal pro b-type natriuretic peptide (NT-proBNP) levels [84]. This may indicate an adaptive MnSOD response to cardiac stress rather than direct pathogenic involvement. Recent findings showed that heat stress downregulates MnSOD expression in endothelial cells via p53-mediated disruption of Sp1 binding to the MnSOD promoter, resulting in oxidative damage; this suppression was rescued by treatment with the mitochondria-targeted antioxidant MitoQ10 or the p53 inhibitor Pifithrin-α, highlighting potential therapeutic strategies to preserve vascular redox balance under stress conditions [104].
Altogether, MnSOD is essential for cardiovascular health, mitigating oxidative stress in the heart, blood vessels, and circulating cells [83,85,86]. Its deficiency has been linked to hypertension, atherosclerosis, diabetic heart disease, and cardiomyopathy, while its overexpression or adaptive increases can signal either protection or underlying damage depending on the context [84,92,94,96]. Understanding the tissue-specific and disease-dependent roles of MnSOD could help optimize therapeutic strategies, especially those involving MnSOD mimetics, targeted antioxidants like MitoQ10, or gene therapy approaches to modulate oxidative stress [78,100,101,104].

2.7. MnSOD in Cancer

MnSOD plays a multifaceted role in cancer, exerting both tumor-suppressive and tumor-promoting effects depending on the context. As a key mitochondrial antioxidant enzyme, it protects critical components of cellular metabolism, including enzymes in the TCA cycle, from oxidative damage—a function essential for maintaining metabolic integrity and preventing malignant transformation [105,106,107,108,109]. Loss of MnSOD has been shown to enhance tumor growth, invasiveness, and tumorigenicity, while its reintroduction suppresses these features in multiple xenograft models [110,111,112,113,114,115]. Paradoxically, elevated MnSOD expression is sometimes observed in aggressive cancers, where it contributes to resistance to chemotherapy and radiotherapy [7,116,117,118,119,120].
This context-dependent behavior has been demonstrated in clinical data. In diabetic patients with endometrial cancer, high MnSOD levels are linked to worse overall survival, although this effect can be mitigated by statin therapy [121]. Similarly, in colon adenocarcinoma, increased MnSOD expression in adjacent normal tissues—but not within tumors themselves—correlates with poor prognosis, further supporting its role as a nuanced prognostic indicator [122]. In pancreatic cancer, MnSOD is upregulated via peroxisome proliferator-activated receptor gamma (PPARγ), which promotes tumor proliferation by suppressing autophagy-related protein 4D (ATG4D)-mediated mitophagy, thus preventing mitochondrial reactive oxygen species (ROS)-induced apoptosis [123]. This reflects a mechanistic switch whereby MnSOD supports tumor survival under oxidative stress by inhibiting cell death pathways.
MnSOD is encoded by the superoxide dismutase 2 (SOD2) gene on chromosome 6q25.3. While this region is generally stable in cancer, some cases—such as melanoma—exhibit loss of heterozygosity or decreased expression in virally transformed fibroblasts [124,125,126]. Reduced expression often arises from epigenetic silencing, including promoter hypermethylation and chromatin remodeling, whereas overexpression tends to result from transcriptional activation by oxidative stress or inflammatory signaling [124]. Transcription factors such as specificity protein 1 (Sp1) and AP-2 regulate SOD2 transcription, and silencing of AP-2 by hypermethylation has been linked to MnSOD upregulation in aggressive breast cancer [34,127,128]. Mutations in the promoter region that impair AP-2 binding or interfere with Sp1 recruitment further reduce MnSOD transcription [129,130]. These alterations are commonly observed in breast, multiple myeloma, and pancreatic cancer cell lines [131,132,133,134].
The tumor suppressor TP53, encoding the p53 protein, also modulates MnSOD levels. High p53 activity represses MnSOD, while low expression or mutated p53 allows for activation of nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB)-driven transcription [110,135,136,137,138,139]. Notably, MnSOD can physically interact with p53 within mitochondria to initiate apoptosis [140]. This dual regulatory role is evident in gastric cancer, where exogenous thermostable MnSOD inhibits tumor progression, upregulates p53, and downregulates zinc finger E-box-binding homeobox 1, implicating MnSOD in both transcriptional and redox-based tumor suppression [141].
MnSOD transcription is also regulated by stress response pathways. NF-κB binds an intronic enhancer in SOD2, promoting transcriptional activation in metastatic breast cancer [33,34,142], thereby enhancing survival and resistance to apoptosis [143,144]. Tumor necrosis factor alpha-induced NF-κB activation also boosts MnSOD expression to counter ROS toxicity [145]. Similarly, Nrf2 upregulates MnSOD by binding the antioxidant response element upon kelch-like ECH-associated protein 1 (Keap1) inhibition [146,147,148,149,150]. Nrf2 activation or Keap1 mutations can contribute to chemoresistance and MnSOD overexpression [151,152,153,154].
Forkhead box O3a (FoxO3a), another key transcription factor, is modulated by the phosphoinositide 3-kinase (PI3K)/protein kinase B (Akt) pathway. Akt phosphorylates FoxO3a, excluding it from the nucleus and suppressing MnSOD expression [155,156,157]. Sirtuin 3 (SIRT3) enhances FoxO3a transcriptional activity via deacetylation, promoting MnSOD expression during oxidative stress [158]. These interactions are functionally important in glioma, where 6-gingerol induces apoptosis via MnSOD and extracellular signal-regulated kinase (ERK) activation. Knockdown of MnSOD or inhibition of ERK abolishes these effects, demonstrating MnSOD’s central role in stress-mediated cell death [159]. Dimethyl fumarate (DMF) also suppresses tumor progression in oral squamous carcinoma by upregulating MnSOD, reducing oxidative stress, and reversing epithelial–mesenchymal transition [160]. Likewise, a triple drug combination—metformin, efavirenz, and fluoxetine—induces ROS overload and upregulates MnSOD, overwhelming antioxidant defenses and causing selective cancer cell apoptosis [161].
Additional regulators include aryl hydrocarbon receptor nuclear translocator in acute myeloid leukemia (AML), CREB-1, activating transcription factor 1 (ATF-1), and HIF-1α [162,163]. Stem cell-associated factors Nanog and Oct4 can also control MnSOD transcription [164,165,166,167].
Post-transcriptionally, MnSOD is suppressed by microRNAs (miRNAs) targeting its 3′ untranslated region. An Alu-like element within the 3′UTR recruits small RNAs that block translation [168]. For instance, miR-222 and miR-382 downregulate MnSOD in tongue cancer and under transforming growth factor beta signaling [169,170], while low miR-17 and miR-146a levels are associated with increased MnSOD and altered treatment responses in prostate and ovarian cancer [171,172,173].
MnSOD function is also modulated by genetic and post-translational changes. Single-nucleotide polymorphisms like Ile58Thr, Leu60Phe, and Ala16Val reduce activity and have been linked to cancer susceptibility [8,68,174,175,176,177,178,179,180]. In gastric cancer, Val/Val genotypes are enriched, supporting its role as a risk factor [181]. Post-translational modifications such as acetylation, phosphorylation, oxidation, nitration, and ubiquitination affect MnSOD function [124]. Notably, acetylation at lysine 68 (K68) converts MnSOD from an antioxidant tetramer to a pro-oxidant monomer, promoting tumor growth [182]. K68-acetylated MnSOD confers resistance to chemotherapy and enhances the tumorigenic potential, while also promoting stemness through HIF-2α stabilization and activation of Oct4, Sox2, and Nanog [182,183].
MnSOD is deeply integrated into cancer metabolism. It supports the Warburg effect via AMP-activated protein kinase (AMPK) and glucose transporter 1 activation [184,185,186,187,188,189,190]. In gastric cancer, toll-like receptor 2 induces MnSOD and glycolysis, suggesting that MnSOD has potential as a biomarker [191]. MnSOD also supports mitochondrial respiration under nutrient stress and promotes proteasomal degradation, thereby enhancing cancer cell survival during metabolic challenges [192].
Lastly, MnSOD modulates the tumor immune microenvironment. It is associated with C-X-C motif chemokine ligand 8 (CXCL8)-mediated neutrophil infiltration [193], CD68-positive macrophages in lung cancer [194], and M2 macrophage polarization in triple-negative breast cancer [195]. Notably, MnSOD-overexpressing oncolytic viruses recruit lymphocytes and enhance immunotherapy responses [196].
In summary, MnSOD shapes nearly every facet of tumor biology—from redox balance and metabolism to immune evasion and treatment resistance. Its dual roles, governed by diverse transcriptional, post-transcriptional, and post-translational mechanisms, highlight its potential as both a therapeutic target and diagnostic biomarker in oncology.

2.8. Implication in Neurodegenerative Diseases

Manganese superoxide dismutase (MnSOD) is primarily expressed in neurons throughout the brain and spinal cord, with lower basal levels in astrocytes; however, its expression in astrocytes increases significantly under conditions of oxidative stress [197]. Brain-specific SOD2-deficient transgenic mice exhibit perinatal death, a lower body weight, a reduced brain size, spongiform encephalopathy, and increased astrocyte activation in the cerebral cortex, brain stem, and hippocampus, alongside selective loss of mitochondrial complex II activity and elevated lipid peroxidation, consistent with oxidative stress-induced brain damage [197]. Epilepsy patients show increased reactive oxygen species (ROS) levels and decreased expression of antioxidant enzymes, including MnSOD [198]. Lower MnSOD expression is associated with greater oxidative damage and heightened neuronal excitability, both of which contribute to seizure susceptibility [198]. Preclinical models indicate that MnSOD mimetics alleviate oxidative stress and reduce seizure severity [199]. Clinical findings also suggest that apoptotic markers are elevated in epilepsy patients, particularly in those with the MnSOD Ala16Val polymorphism. This polymorphism is associated with increased interleukin-1β (IL-1β) and interleukin-6 (IL-6) levels, implicating MnSOD in inflammation-linked seizure pathology and highlighting the relevance of genetic variation in disease severity and immune responses [200]. Huntington’s disease is characterized by oxidative stress and mitochondrial dysfunction caused by mutant huntingtin, which disrupts mitochondrial respiration and elevates ROS levels. MnSOD helps attenuate ROS accumulation and protect neurons, with overexpression of MnSOD conferring neuroprotection in disease models [201,202]. In a 3-nitropropionic acid-induced model of Huntington’s disease, treatment with C60 fullerene reduced oxidative stress, improved mitochondrial function, and restored MnSOD levels, demonstrating its therapeutic potential through antioxidant pathway modulation [203].
Oxidative stress is also a major contributor to ischemic stroke, where MnSOD levels are often reduced in affected patients [204]. In transgenic mice, MnSOD overexpression or treatment with MnSOD mimetics mitigates ischemia/reperfusion injury and improves neurological outcomes [204,205]. Administration of irisin, a myokine linked to physical exercise, in a murine model of cerebral ischemia reduced mortality and improved cognition, effects that were associated with increased expression of MnSOD and the anti-aging protein Klotho, suggesting that MnSOD mediates the neuroprotective benefits of exercise-mimetic therapies in stroke [206]. Similarly, chronic low-dose alcohol consumption in a post-ischemic model attenuated inflammation via activation of peroxisome proliferator-activated receptor gamma (PPARγ), which correlated with increased MnSOD expression and reduced neuroinflammation [207]. Clinical data show that MnSOD levels inversely correlate with stroke severity, suggesting its potential as both a biomarker and therapeutic target [204,205]. In a rodent model of stroke, treatment with chlorpromazine and promethazine reduced infarct size and neuronal damage through modulation of the protein kinase C delta (PKC-δ)/nicotinamide adenine dinucleotide phosphate oxidase/MnSOD pathway, confirming MnSOD’s involvement in pharmacological neuroprotection during ischemic injury [208]. Notoginseng leaf triterpenes (NLTs), a compound derived from Panax notoginseng, protected against cerebral ischemia and oxygen-glucose deprivation/reoxygenation-induced neuronal injury by activating the nicotinamide phosphoribosyltransferase–sirtuin (SIRT)1/2/3–Forkhead box O3 (Foxo3a)–MnSOD signaling pathway, demonstrating MnSOD’s role as a key effector in NAD+-dependent antioxidant defenses [209,210].
In Parkinson’s disease (PD), oxidative stress contributes to dopaminergic neuron loss in the substantia nigra. Overexpression of MnSOD in transgenic mice improves survival following neurotoxin exposure, while MnSOD mimetics reduce ROS and neuroinflammation [211,212]. The mitochondrial deacetylase SIRT3 enhances MnSOD activity by preserving its active tetrameric structure through deacetylation, thereby reducing mitochondrial ROS and offering a neuroprotective mechanism in PD [213]. Additionally, inhibition of mixed-lineage kinase domain-like protein in a 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-induced PD mouse model decreased oxidative stress and dopaminergic neuronal death, in part by increasing MnSOD expression, further implicating MnSOD in limiting PD-associated neurodegeneration [214]. Clinical studies have reported decreased MnSOD levels in patients, linking oxidative damage to disease progression [212].
In Alzheimer’s disease (AD), MnSOD plays a similar neuroprotective role. Oxidative stress and mitochondrial dysfunction are central features of AD, and MnSOD overexpression in transgenic mice reduces hippocampal superoxide levels and alleviates memory deficits [215,216]. In multiple AD models, modulation of MnSOD expression resulted in consistent neuroprotection, with reductions in ROS and improvements in cognitive function [216]. In a triple-transgenic AD (3×Tg-AD) mouse model with cerebral microinfarcts, activation of protein kinase C epsilon (PKCε) improved memory and hippocampal vascular integrity, effects associated with increased MnSOD expression, suggesting that MnSOD contributes to both neurovascular and cognitive protection in AD complicated by vascular injury [217]. In aged and AD hippocampi, PKCε activation also restored PKCε, MnSOD, and vascular endothelial growth factor levels and microvessel density, implicating MnSOD in a PKCε-regulated neurovascular repair axis that becomes dysregulated in aging and AD [218]. MnSOD is also significantly upregulated by DMF, a treatment used in neurodegeneration models, leading to improved mitochondrial function and redox balance. These effects are abrogated by silencing Nrf2, confirming MnSOD as a key downstream target of Nrf2-driven neuroprotection [219]. MnSOD expression is also increased via SIRT3 activation in amyloid-beta (Aβ)-treated models, where a novel rhamnoside derivative reduced ROS, restored mitochondrial integrity, and suppressed cellular senescence, reinforcing the MnSOD–SIRT3 axis as a promising neuroprotective pathway in AD [220]. Pramlintide, an amylin analog, enhanced cortical MnSOD expression and mitochondrial function while reducing oxidative damage in AD models, further validating MnSOD’s protective role [221]. In apolipoprotein E4 (ApoE4)-positive AD models, MnSOD expression in hippocampal microvessels was reduced, implicating MnSOD in vascular oxidative stress and suggesting a role for MnSOD in genetically predisposed cognitive decline [222]. Consistent with this, irradiated hippocampal slices from ApoE4 mice exhibited impaired ROS responses due to diminished MnSOD expression, underscoring MnSOD’s role in adaptive redox signaling under environmental and genetic stress [223].
Beyond classical neurodegenerative diseases, MnSOD also plays a protective role in neuroinflammatory disorders. In multiple sclerosis (MS), MnSOD expression is upregulated following ocrelizumab treatment, particularly in CD19+ B cells, suggesting a role in the drug’s neuroprotective and anti-inflammatory effects through PKCβ modulation [224]. Vitamin D supplementation in patients with relapsing–remitting MS also significantly increases MnSOD expression, supporting its function in broader antioxidant defenses during neuroinflammation [225]. Mice with neuron-specific MnSOD deletion exhibit demyelination, inflammation, and progressive paralysis, recapitulating the pathological features of progressive MS and demonstrating the importance of MnSOD in maintaining neuronal resilience against inflammation-induced damage [226]. Similarly, in experimental autoimmune encephalomyelitis, combination therapy with lovastatin and an AMPK activator improved mitochondrial and peroxisomal function, upregulated MnSOD, and reduced disease severity, positioning MnSOD as a therapeutic effector in metabolically guided treatment of MS [227].
In summary, MnSOD is vital for neuronal health throughout the central nervous system. It is constitutively expressed in neurons and becomes upregulated in astrocytes during oxidative stress [197]. Genetic deletion results in profound neurological deficits and early mortality [197], while reduced MnSOD expression is commonly observed in conditions such as epilepsy, Huntington’s disease, ischemic stroke, Parkinson’s disease, and Alzheimer’s disease [198,202,204,211,215]. In each of these disorders, therapeutic elevation of MnSOD levels—through genetic overexpression, pharmacologic mimetics, or upstream regulators like SIRT3 or Nrf2—consistently reduces the ROS burden and limits neuronal injury [199,201,205,212,216]. Therefore, MnSOD represents a promising target for managing oxidative stress for a broad spectrum of neurodegenerative and neuroinflammatory diseases.

3. Conclusions

MnSOD is a vital mitochondrial antioxidant enzyme whose structural stability, catalytic activity, and regulatory flexibility are central to maintaining cellular redox homeostasis. Its role in converting superoxide into H2O2 makes it essential for protecting mitochondrial function while also influencing broader metabolic and signaling pathways. MnSOD’s physiological importance spans thermogenesis, energy metabolism, and neuronal health, while its dysregulation is implicated in the development and progression of obesity, kidney disease, cardiovascular dysfunction, cancer, and neurodegeneration. Obesity impairs MnSOD expression and mitochondrial efficiency in brown adipose tissue and skeletal muscle, contributing to thermogenic failure and systemic metabolic dysfunction. Loss of MnSOD activity exacerbates lipid accumulation and insulin resistance, while its preservation supports mitochondrial respiration and energy expenditure. In the kidney, MnSOD dysfunction drives oxidative injury across multiple nephron compartments. Whether in acute kidney injury, chronic allograft rejection, or diabetic nephropathy, its loss results in elevated mitochondrial ROS, inflammation, and fibrosis. Restoration of MnSOD through genetic or pharmacologic means confers significant protection. The cardiovascular effects of MnSOD are highly context- and tissue-dependent: deficiency promotes hypertension, atherosclerosis, cardiomyopathy, and endothelial dysfunction, whereas upregulation may indicate either protective adaptation or pathological stress signaling. Within cancer biology, MnSOD assumes a dual role—suppressing tumor initiation by preserving mitochondrial integrity and limiting DNA damage, but later promoting tumor progression, angiogenesis, and therapy resistance through altered regulation and metabolic reprogramming. Finally, in neurodegenerative disorders, MnSOD deficiency heightens neuronal vulnerability to mitochondrial dysfunction and oxidative damage. This is particularly evident in Alzheimer’s, Parkinson’s, and diabetic retinopathy models, where its maintenance improves mitochondrial integrity, redox balance, and neuronal survival. These context-dependent effects are shaped by a network of regulatory mechanisms, including transcription factors, epigenetic marks, microRNAs, and post-translational changes that modulate MnSOD expression, localization, and activity. In several disease models, enhancing MnSOD levels or function, whether through genetic means, transcriptional activation, or targeted post-translational modulation, confers protective effects, while its deficiency exacerbates oxidative damage and dysfunction. To complement this, we have published a 2024 review of the current state of MnSOD mimetics, which have shown promise as pharmacological tools to replicate or enhance MnSOD activity in pathological settings [228]. Continued research into natural MnSOD regulation and the rational development of mimetics may yield effective strategies for managing oxidative stress-related diseases, particularly those in which direct genetic modulation is impractical. Together, these complementary approaches hold significant promise for advancing the treatment of a wide range of oxidative stress-driven conditions.

Author Contributions

Conceptualization, J.G. and A.R.A.; writing—original draft preparation, J.G. and A.R.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the NIH/NCI under Grant 1R01CA258673. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript (ARA).

Acknowledgments

We thank Kerry Evans and the UAMS Science Communication Group for editing this manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ADAlzheimer’s disease
AFAtrial fibrillation
AMLAcute myeloid leukemia
AMPKAMP-activated protein kinase
AP-1
AP-2
Activator protein 1
Activator protein 2
ATFActivating transcription factor
CPETConcerted proton–electron transfer
CREBcAMP response element-binding protein
DJ-1the multifunctional redox-sensitive protein DJ-1 (PARK7)
DMFDimethyl fumarate
ERKExtracellular signal-regulated kinase
HIF-1αHypoxia-inducible factor 1-alpha
ILInterleukin
MSMultiple sclerosis
NADNicotinamide adenine dinucleotide
NADHReduced nicotinamide adenine dinucleotide
NF-κBNuclear factor kappa B
NRF2Nuclear factor erythroid 2-related factor 2
ONOOPeroxynitrite
PDParkinson’s disease
PGC-1αPeroxisome proliferator-activated receptor gamma coactivator 1-alpha
PKCProtein kinase C
PPHNPersistent pulmonary hypertension of the newborn
ROSReactive oxygen species
SIRT3Sirtuin 3
STZStreptozotocin
TCATricarboxylic acid
WAT1First-shell water ligand bound to manganese ion

References

  1. San-Millán, I. The Key Role of Mitochondrial Function in Health and Disease. Antioxidants 2023, 12, 782. [Google Scholar] [CrossRef] [PubMed]
  2. Dan Dunn, J.; Alvarez, L.A.J.; Zhang, X.; Soldati, T. Reactive oxygen species and mitochondria: A nexus of cellular homeostasis. Redox Biol. 2015, 6, 472–485. [Google Scholar] [CrossRef]
  3. Zheng, M.; Liu, Y.; Zhang, G.; Yang, Z.; Xu, W.; Chen, Q. The Applications and Mechanisms of Superoxide Dismutase in Medicine, Food, and Cosmetics. Antioxidants 2023, 12, 1675. [Google Scholar] [CrossRef] [PubMed]
  4. Younus, H. Therapeutic potentials of superoxide dismutase. Int. J. Health Sci. 2018, 12, 88–93. [Google Scholar]
  5. Mondola, P.; Damiano, S.; Sasso, A.; Santillo, M. The Cu, Zn Superoxide Dismutase: Not Only a Dismutase Enzyme. Front. Physiol. 2016, 7, 594. [Google Scholar] [CrossRef] [PubMed]
  6. Mishra, N.; Jiang, C.; Chen, L.; Paul, A.; Chatterjee, A.; Shen, G. Achieving abiotic stress tolerance in plants through antioxidative defense mechanisms. Front. Plant Sci. 2023, 14, 1110622. [Google Scholar] [CrossRef]
  7. Liu, M.; Sun, X.; Chen, B.; Dai, R.; Xi, Z.; Xu, H. Insights into Manganese Superoxide Dismutase and Human Diseases. Int. J. Mol. Sci. 2022, 23, 15893. [Google Scholar] [CrossRef]
  8. Borgstahl, G.E.; Parge, H.E.; Hickey, M.J.; Beyer, W.F., Jr.; Hallewell, R.A.; Tainer, J.A. The structure of human mitochondrial manganese superoxide dismutase reveals a novel tetrameric interface of two 4-helix bundles. Cell 1992, 71, 107–118. [Google Scholar] [CrossRef]
  9. Karnati, S.; Lüers, G.; Pfreimer, S.; Baumgart-Vogt, E. Mammalian SOD2 is exclusively located in mitochondria and not present in peroxisomes. Histochem. Cell Biol. 2013, 140, 105–117. [Google Scholar] [CrossRef]
  10. McCord, J.M.; Fridovich, I. The reduction of cytochrome c by milk xanthine oxidase. J. Biol. Chem. 1968, 243, 5753–5760. [Google Scholar] [CrossRef]
  11. McCord, J.M.; Fridovich, I. Superoxide dismutase. An enzymic function for erythrocuprein (hemocuprein). J. Biol. Chem. 1969, 244, 6049–6055. [Google Scholar] [CrossRef] [PubMed]
  12. Keele, B.B.; McCord, J.; Fridovich, I. Superoxide dismutase from Escherichia coli B: A new manganese-containing enzyme. J. Biol. Chem. 1970, 245, 6176–6181. [Google Scholar] [CrossRef] [PubMed]
  13. Creagan, R.; Tischfield, J.; Ricciuti, F.; Ruddle, F.H. Chromosome assignments of genes in man using mouse-human somatic cell hybrids: Mitochondrial superoxide dismutase (indophenol oxidase-B, tetrameric) to chromosome 6. Humangenetik 1973, 20, 203–209. [Google Scholar] [CrossRef] [PubMed]
  14. Barra, D.; Schinina, M.E.; Simmaco, M.; Bannister, J.V.; Bannister, W.H.; Rotilio, G.; Bossa, F. The primary structure of human liver manganese superoxide dismutase. J. Biol. Chem. 1984, 259, 12595–12601. [Google Scholar] [CrossRef]
  15. Chidambaram, S.B.; Anand, N.; Varma, S.R.; Ramamurthy, S.; Vichitra, C.; Sharma, A.; Mahalakshmi, A.M.; Essa, M.M. Superoxide dismutase and neurological disorders. IBRO Neurosci. Rep. 2024, 16, 373–394. [Google Scholar] [CrossRef]
  16. Kitada, M.; Xu, J.; Ogura, Y.; Monno, I.; Koya, D. Manganese Superoxide Dismutase Dysfunction and the Pathogenesis of Kidney Disease. Front. Physiol. 2020, 11, 755. [Google Scholar] [CrossRef]
  17. Ninić, A.; Sopić, M.; Munjas, J.; Spasojević-Kalimanovska, V.; Kotur-Stevuljević, J.; Bogavac-Stanojević, N.; Ivanišević, J.; Simić-Ogrizović, S.; Kravljača, M.; Jelić-Ivanović, Z. Association Between Superoxide Dismutase Isoenzyme Gene Expression and Total Antioxidant Status in Patients with an End-Stage Renal Disease. Balkan. Med. J. 2018, 35, 431–436. [Google Scholar] [CrossRef]
  18. Case, A.J. On the Origin of Superoxide Dismutase: An Evolutionary Perspective of Superoxide-Mediated Redox Signaling. Antioxidants 2017, 6, 82. [Google Scholar] [CrossRef]
  19. Quint, P.S.; Domsic, J.F.; Cabelli, D.E.; McKenna, R.; Silverman, D.N. Role of a glutamate bridge spanning the dimeric interface of human manganese superoxide dismutase. Biochemistry 2008, 47, 4621–4628. [Google Scholar] [CrossRef]
  20. Bonetta Valentino, R. The structure-function relationships and physiological roles of MnSOD mutants. Biosci. Rep. 2022, 42, BSR20220202. [Google Scholar] [CrossRef]
  21. Sheng, Y.; Butler Gralla, E.; Schumacher, M.; Cascio, D.; Cabelli, D.E.; Valentine, J.S. Six-coordinate manganese(3+) in catalysis by yeast manganese superoxide dismutase. Proc. Natl. Acad. Sci. USA 2012, 109, 14314–14319. [Google Scholar] [CrossRef] [PubMed]
  22. Hearn, A.S.; Stroupe, M.E.; Cabelli, D.E.; Lepock, J.R.; Tainer, J.A.; Nick, H.S.; Silverman, D.N. Kinetic Analysis of Product Inhibition in Human Manganese Superoxide Dismutase. Biochemistry 2001, 40, 12051–12058. [Google Scholar] [CrossRef]
  23. Azadmanesh, J.; Trickel, S.R.; Borgstahl, G.E.O. Substrate-analog binding and electrostatic surfaces of human manganese superoxide dismutase. J. Struct. Biol. 2017, 199, 68–75. [Google Scholar] [CrossRef]
  24. Greenleaf, W.B.; Perry, J.J.; Hearn, A.S.; Cabelli, D.E.; Lepock, J.R.; Stroupe, M.E.; Tainer, J.A.; Nick, H.S.; Silverman, D.N. Role of hydrogen bonding in the active site of human manganese superoxide dismutase. Biochemistry 2004, 43, 7038–7045. [Google Scholar] [CrossRef]
  25. Azadmanesh, J.; Lutz, W.E.; Coates, L.; Weiss, K.L.; Borgstahl, G.E.O. Direct detection of coupled proton and electron transfers in human manganese superoxide dismutase. Nat. Commun. 2021, 12, 2079. [Google Scholar] [CrossRef] [PubMed]
  26. McAdam, M.E.; Fox, R.A.; Lavelle, F.; Fielden, E.M. A pulse-radiolysis study of the manganese-containing superoxide dismutase from Bacillus stearothermophilus. A kinetic model for the enzyme action. Biochem. J. 1977, 165, 71–79. [Google Scholar] [CrossRef] [PubMed]
  27. Abreu, I.A.; Rodriguez, J.A.; Cabelli, D.E. Theoretical studies of manganese and iron superoxide dismutases: Superoxide binding and superoxide oxidation. J. Phys. Chem. B 2005, 109, 24502–24509. [Google Scholar] [CrossRef]
  28. Perry, J.J.; Hearn, A.S.; Cabelli, D.E.; Nick, H.S.; Tainer, J.A.; Silverman, D.N. Contribution of human manganese superoxide dismutase tyrosine 34 to structure and catalysis. Biochemistry 2009, 48, 3417–3424. [Google Scholar] [CrossRef]
  29. Azadmanesh, J.; Slobodnik, K.; Struble, L.R.; Lutz, W.E.; Coates, L.; Weiss, K.L.; Myles, D.A.A.; Kroll, T.; Borgstahl, G.E.O. Revealing the atomic and electronic mechanism of human manganese superoxide dismutase product inhibition. Nat. Commun. 2024, 15, 5973. [Google Scholar] [CrossRef]
  30. Zhang, X.; Liu, H.; Zhang, D. MnSOD serves as the central molecule in adaptive thermogenesis (MnSOD functions as a thermoreceptor). Adv. Redox Res. 2021, 3, 100027. [Google Scholar] [CrossRef]
  31. Xu, Y.; Porntadavity, S.; St Clair, D.K. Transcriptional regulation of the human manganese superoxide dismutase gene: The role of specificity protein 1 (Sp1) and activating protein-2 (AP-2). Biochem. J. 2002, 362, 401–412. [Google Scholar] [CrossRef] [PubMed]
  32. Cherry, A.D.; Suliman, H.B.; Bartz, R.R.; Piantadosi, C.A. Peroxisome proliferator-activated receptor γ co-activator 1-α as a critical co-activator of the murine hepatic oxidative stress response and mitochondrial biogenesis in Staphylococcus aureus sepsis. J. Biol. Chem. 2014, 289, 41–52. [Google Scholar] [CrossRef]
  33. Jones, P.L.; Ping, D.; Boss, J.M. Tumor necrosis factor alpha and interleukin-1beta regulate the murine manganese superoxide dismutase gene through a complex intronic enhancer involving C/EBP-beta and NF-kappaB. Mol. Cell Biol. 1997, 17, 6970–6981. [Google Scholar] [CrossRef]
  34. Ennen, M.; Minig, V.; Grandemange, S.; Touche, N.; Merlin, J.L.; Besancenot, V.; Brunner, E.; Domenjoud, L.; Becuwe, P. Regulation of the high basal expression of the manganese superoxide dismutase gene in aggressive breast cancer cells. Free Radic. Biol. Med. 2011, 50, 1771–1779. [Google Scholar] [CrossRef]
  35. Tao, R.; Coleman, M.C.; Pennington, J.D.; Ozden, O.; Park, S.H.; Jiang, H.; Kim, H.S.; Flynn, C.R.; Hill, S.; Hayes McDonald, W.; et al. Sirt3-mediated deacetylation of evolutionarily conserved lysine 122 regulates MnSOD activity in response to stress. Mol. Cell 2010, 40, 893–904. [Google Scholar] [CrossRef] [PubMed]
  36. Chen, Y.; Zhang, J.; Lin, Y.; Lei, Q.; Guan, K.L.; Zhao, S.; Xiong, Y. Tumour suppressor SIRT3 deacetylates and activates manganese superoxide dismutase to scavenge ROS. EMBO Rep. 2011, 12, 534–541. [Google Scholar] [CrossRef] [PubMed]
  37. MacMillan-Crow, L.A.; Crow, J.P.; Thompson, J.A. Peroxynitrite-mediated inactivation of manganese superoxide dismutase involves nitration and oxidation of critical tyrosine residues. Biochemistry 1998, 37, 1613–1622. [Google Scholar] [CrossRef]
  38. Hsu, J.-L.; Hsieh, Y.; Tu, C.; O’Connor, D.; Nick, H.S.; Silverman, D.N. Catalytic Properties of Human Manganese Superoxide Dismutase*. J. Biol. Chem. 1996, 271, 17687–17691. [Google Scholar] [CrossRef]
  39. Buettner, G.R.; Ng, C.F.; Wang, M.; Rodgers, V.G.J.; Schafer, F.Q. A New Paradigm: Manganese Superoxide Dismutase Influences the Production of H2O2 in Cells and Thereby Their Biological State. Free Radic. Biol. Med. 2006, 41, 1338–1350. [Google Scholar] [CrossRef]
  40. Chouchani, E.T.; Kazak, L.; Jedrychowski, M.P.; Lu, G.Z.; Erickson, B.K.; Szpyt, J.; Pierce, K.A.; Laznik-Bogoslavski, D.; Vetrivelan, R.; Clish, C.B.; et al. Mitochondrial ROS regulate thermogenic energy expenditure and sulfenylation of UCP1. Nature 2016, 532, 112–116. [Google Scholar] [CrossRef]
  41. Ježek, J.; Dlasková, A.; Zelenka, J.; Jabůrek, M.; Ježek, P. H2O2-Activated Mitochondrial Phospholipase iPLA2γ Prevents Lipotoxic Oxidative Stress in Synergy with UCP2, Amplifies Signaling via G-Protein–Coupled Receptor GPR40, and Regulates Insulin Secretion in Pancreatic β-Cells. Antioxid. Redox Signal. 2015, 23, 958–972. [Google Scholar] [CrossRef] [PubMed]
  42. Jabůrek, M.; Ježek, J.; Zelenka, J.; Ježek, P. Antioxidant activity by a synergy of redox-sensitive mitochondrial phospholipase A2 and uncoupling protein-2 in lung and spleen. Int. J. Biochem. Cell Biol. 2013, 45, 816–825. [Google Scholar] [CrossRef]
  43. Yamamoto, S.; Shimizu, S. Significance of TRP channels in oxidative stress. Eur. J. Pharmacol. 2016, 793, 109–111. [Google Scholar] [CrossRef] [PubMed]
  44. Sakaguchi, R.; Mori, Y. Transient receptor potential (TRP) channels: Biosensors for redox environmental stimuli and cellular status. Free Radic. Biol. Med. 2020, 146, 36–44. [Google Scholar] [CrossRef] [PubMed]
  45. Han, Y.H.; Buffolo, M.; Pires, K.M.; Pei, S.; Scherer, P.E.; Boudina, S. Adipocyte-Specific Deletion of Manganese Superoxide Dismutase Protects From Diet-Induced Obesity Through Increased Mitochondrial Uncoupling and Biogenesis. Diabetes 2016, 65, 2639–2651. [Google Scholar] [CrossRef]
  46. Montano, M.A.E.; Barrio Lera, J.P.; Gottlieb, M.G.V.; Schwanke, C.H.A.; da Rocha, M.I.U.M.; Manica-Cattani, M.F.; dos Santos, G.F.; da Cruz, I.B.M. Association between manganese superoxide dismutase (MnSOD) gene polymorphism and elderly obesity. Mol. Cell. Biochem. 2009, 328, 33–40. [Google Scholar] [CrossRef]
  47. Garcia-Irigoyen, O.; Bovenga, F.; Piglionica, M.; Piccinin, E.; Cariello, M.; Arconzo, M.; Peres, C.; Corsetto, P.A.; Rizzo, A.M.; Ballanti, M.; et al. Enterocyte superoxide dismutase 2 deletion drives obesity. iScience 2022, 25, 103707. [Google Scholar] [CrossRef]
  48. Boyle, K.E.; Newsom, S.A.; Janssen, R.C.; Lappas, M.; Friedman, J.E. Skeletal Muscle MnSOD, Mitochondrial Complex II, and SIRT3 Enzyme Activities Are Decreased in Maternal Obesity During Human Pregnancy and Gestational Diabetes Mellitus. J. Clin. Endocrinol. Metab. 2013, 98, E1601–E1609. [Google Scholar] [CrossRef]
  49. Che, R.; Yuan, Y.; Huang, S.; Zhang, A. Mitochondrial dysfunction in the pathophysiology of renal diseases. Am. J. Physiol. Renal. Physiol. 2014, 306, F367–F378. [Google Scholar] [CrossRef]
  50. He, L.; Wei, Q.; Liu, J.; Yi, M.; Liu, Y.; Liu, H.; Sun, L.; Peng, Y.; Liu, F.; Venkatachalam, M.A.; et al. AKI on CKD: Heightened injury, suppressed repair, and the underlying mechanisms. Kidney Int. 2017, 92, 1071–1083. [Google Scholar] [CrossRef]
  51. Rodriguez-Iturbe, B.; Sepassi, L.; Quiroz, Y.; Ni, Z.; Wallace, D.C.; Vaziri, N.D. Association of mitochondrial SOD deficiency with salt-sensitive hypertension and accelerated renal senescence. J. Appl. Physiol. 2007, 102, 255–260. [Google Scholar] [CrossRef] [PubMed]
  52. Noiri, E.; Nakao, A.; Uchida, K.; Tsukahara, H.; Ohno, M.; Fujita, T.; Brodsky, S.; Goligorsky, M.S. Oxidative and nitrosative stress in acute renal ischemia. Am. J. Physiol. Renal. Physiol. 2001, 281, F948–F957. [Google Scholar] [CrossRef] [PubMed]
  53. Ouyang, J.; Zeng, Z.; Fang, H.; Li, F.; Zhang, X.; Tan, W. SIRT3 Inactivation Promotes Acute Kidney Injury Through Elevated Acetylation of SOD2 and p53. J. Surg. Res. 2019, 233, 221–230. [Google Scholar] [CrossRef] [PubMed]
  54. Parajuli, N.; Marine, A.; Simmons, S.; Saba, H.; Mitchell, T.; Shimizu, T.; Shirasawa, T.; Macmillan-Crow, L.A. Generation and characterization of a novel kidney-specific manganese superoxide dismutase knockout mouse. Free Radic. Biol. Med. 2011, 51, 406–416. [Google Scholar] [CrossRef]
  55. MacMillan-Crow, L.A.; Crow, J.P.; Kerby, J.D.; Beckman, J.S.; Thompson, J.A. Nitration and inactivation of manganese superoxide dismutase in chronic rejection of human renal allografts. Proc. Natl. Acad. Sci. USA 1996, 93, 11853–11858. [Google Scholar] [CrossRef]
  56. Sharma, K. Mitochondrial Dysfunction in the Diabetic Kidney. Adv. Exp. Med. Biol. 2017, 982, 553–562. [Google Scholar] [CrossRef]
  57. Xu, J.; Kitada, M.; Koya, D. The impact of mitochondrial quality control by Sirtuins on the treatment of type 2 diabetes and diabetic kidney disease. Biochim. Biophys. Acta Mol. Basis. Dis. 2020, 1866, 165756. [Google Scholar] [CrossRef]
  58. Ma, S.X.; Li, X.J.; Duan, T.T.; Pei, M.; Zou, L.; Yu, X.Y.; Zhao, Y.Y. Moshen granule ameliorates membranous nephropathy by regulating NF-ƙB/Nrf2 pathways via aryl hydrocarbon receptor signalling. Heliyon 2023, 9, e20019. [Google Scholar] [CrossRef]
  59. Trambas, I.A.; Bowen, L.; Thallas-Bonke, V.; Snelson, M.; Sourris, K.C.; Laskowski, A.; Tauc, M.; Rubera, I.; Zheng, G.; Harris, D.C.H.; et al. Proximal tubular deletion of superoxide dismutase-2 reveals disparate effects on kidney function in diabetes. Redox. Biol. 2025, 82, 103601. [Google Scholar] [CrossRef]
  60. Shimoda-Matsubayashi, S.; Matsumine, H.; Kobayashi, T.; Nakagawa-Hattori, Y.; Shimizu, Y.; Mizuno, Y. Structural dimorphism in the mitochondrial targeting sequence in the human manganese superoxide dismutase gene. A predictive evidence for conformational change to influence mitochondrial transport and a study of allelic association in Parkinson’s disease. Biochem. Biophys. Res. Commun. 1996, 226, 561–565. [Google Scholar] [CrossRef]
  61. Möllsten, A.; Marklund, S.L.; Wessman, M.; Svensson, M.; Forsblom, C.; Parkkonen, M.; Brismar, K.; Groop, P.H.; Dahlquist, G. A functional polymorphism in the manganese superoxide dismutase gene and diabetic nephropathy. Diabetes 2007, 56, 265–269. [Google Scholar] [CrossRef]
  62. Nomiyama, T.; Tanaka, Y.; Piao, L.; Nagasaka, K.; Sakai, K.; Ogihara, T.; Nakajima, K.; Watada, H.; Kawamori, R. The polymorphism of manganese superoxide dismutase is associated with diabetic nephropathy in Japanese type 2 diabetic patients. J. Hum. Genet. 2003, 48, 138–141. [Google Scholar] [CrossRef]
  63. Houldsworth, A.; Hodgkinson, A.; Shaw, S.; Millward, A.; Demaine, A.G. Polymorphic differences in the SOD-2 gene may affect the pathogenesis of nephropathy in patients with diabetes and diabetic complications. Gene 2015, 569, 41–45. [Google Scholar] [CrossRef] [PubMed]
  64. Lee, S.J.; Choi, M.G.; Kim, D.S.; Kim, T.W. Manganese superoxide dismutase gene polymorphism (V16A) is associated with stages of albuminuria in Korean type 2 diabetic patients. Metabolism 2006, 55, 1–7. [Google Scholar] [CrossRef] [PubMed]
  65. Liu, L.; Zheng, T.; Wang, N.; Wang, F.; Li, M.; Jiang, J.; Zhao, R.; Li, L.; Zhao, W.; Zhu, Q.; et al. The manganese superoxide dismutase Val16Ala polymorphism is associated with decreased risk of diabetic nephropathy in Chinese patients with type 2 diabetes. Mol. Cell Biochem. 2009, 322, 87–91. [Google Scholar] [CrossRef] [PubMed]
  66. Ascencio-Montiel Ide, J.; Parra, E.J.; Valladares-Salgado, A.; Gómez-Zamudio, J.H.; Kumate-Rodriguez, J.; Escobedo-de-la-Peña, J.; Cruz, M. SOD2 gene Val16Ala polymorphism is associated with macroalbuminuria in Mexican type 2 diabetes patients: A comparative study and meta-analysis. BMC Med. Genet. 2013, 14, 110. [Google Scholar] [CrossRef]
  67. Tian, C.; Fang, S.; Du, X.; Jia, C. Association of the C47T polymorphism in SOD2 with diabetes mellitus and diabetic microvascular complications: A meta-analysis. Diabetologia 2011, 54, 803–811. [Google Scholar] [CrossRef]
  68. Sutton, A.; Khoury, H.; Prip-Buus, C.; Cepanec, C.; Pessayre, D.; Degoul, F. The Ala16Val genetic dimorphism modulates the import of human manganese superoxide dismutase into rat liver mitochondria. Pharmacogenetics 2003, 13, 145–157. [Google Scholar] [CrossRef]
  69. Ashcroft, F.M.; Rorsman, P. Diabetes mellitus and the β cell: The last ten years. Cell 2012, 148, 1160–1171. [Google Scholar] [CrossRef]
  70. Rains, J.L.; Jain, S.K. Oxidative stress, insulin signaling, and diabetes. Free Radic. Biol. Med. 2011, 50, 567–575. [Google Scholar] [CrossRef]
  71. Wu, J.; Zhou, Y.; Hu, H.; Yang, D.; Yang, F. Effects of β-carotene on glucose metabolism dysfunction in humans and type 2 diabetic rats. Acta Mater. Medica 2022, 1, 138–153. [Google Scholar] [CrossRef]
  72. Suri, S.; Mitra, P.; Bankul, A.; Saxena, I.; Garg, M.K.; Bohra, G.K.; Sharma, P. Altered expression of specific antioxidant (SOD1 and SOD2) and DNA repair (XRCC1 and OGG1) genes in patients with newly diagnosed type-2 diabetes mellitus. Minerva Endocrinol 2021, 49, 398–405. [Google Scholar] [CrossRef]
  73. Eddaikra, A.; Amroun, H.; Raache, R.; Galleze, A.; Abdallah-Elhadj, N.; Azzouz, M.; Meçabih, F.; Mechti, B.; Abbadi, M.C.; Touil-Boukoffa, C.; et al. Clinical variables and ethnicity may influenced by polymorphism of CAT -262C/T and MnSOD 47C/T antioxidant enzymes in Algerian type1 diabetes without complications. Gene 2018, 670, 182–192. [Google Scholar] [CrossRef] [PubMed]
  74. Chen, H.; Li, X.; Epstein, P.N. MnSOD and catalase transgenes demonstrate that protection of islets from oxidative stress does not alter cytokine toxicity. Diabetes 2005, 54, 1437–1446. [Google Scholar] [CrossRef]
  75. Bertera, S.; Crawford, M.L.; Alexander, A.M.; Papworth, G.D.; Watkins, S.C.; Robbins, P.D.; Trucco, M. Gene transfer of manganese superoxide dismutase extends islet graft function in a mouse model of autoimmune diabetes. Diabetes 2003, 52, 387–393. [Google Scholar] [CrossRef] [PubMed]
  76. Zhang, P.; Li, T.; Wu, X.; Nice, E.C.; Huang, C.; Zhang, Y. Oxidative stress and diabetes: Antioxidative strategies. Front. Med. 2020, 14, 583–600. [Google Scholar] [CrossRef]
  77. Ivanović-Matić, S.; Mihailović, M.; Dinić, S.; Martinović, V.; Bogojević, D.; Grigorov, I.; Poznanović, G. The absence of cardiomyopathy is accompanied by increased activities of CAT, MnSOD and GST in long-term diabetes in rats. J. Physiol. Sci. 2010, 60, 259–266. [Google Scholar] [CrossRef]
  78. Shen, X.; Zheng, S.; Metreveli, N.S.; Epstein, P.N. Protection of Cardiac Mitochondria by Overexpression of MnSOD Reduces Diabetic Cardiomyopathy. Diabetes 2006, 55, 798–805. [Google Scholar] [CrossRef]
  79. Elrashidy, R.A.; Kavran, M.; Asker, M.E.; Mohamed, H.E.; Daneshgari, F.; Liu, G. Smooth muscle-specific deletion of MnSOD exacerbates diabetes-induced bladder dysfunction in mice. Am. J. Physiol. Renal. Physiol. 2019, 317, F906–F912. [Google Scholar] [CrossRef]
  80. Madsen-Bouterse, S.A.; Zhong, Q.; Mohammad, G.; Ho, Y.S.; Kowluru, R.A. Oxidative damage of mitochondrial DNA in diabetes and its protection by manganese superoxide dismutase. Free Radic. Res. 2010, 44, 313–321. [Google Scholar] [CrossRef]
  81. Peng, H.; Li, H.; Ma, B.; Sun, X.; Chen, B. DJ-1 regulates mitochondrial function and promotes retinal ganglion cell survival under high glucose-induced oxidative stress. Front. Pharmacol. 2024, 15, 1455439. [Google Scholar] [CrossRef]
  82. Lv, T.; Lu, Y.; Liu, Y.; Feng, H.; Li, C.; Sheng, W.; Cui, Z.; Zhu, S.; Gu, X.; Yang, Z.; et al. General Control of Amino Acid Synthesis 5-Like 1-Mediated Acetylation of Manganese Superoxide Dismutase Regulates Oxidative Stress in Diabetic Kidney Disease. Oxid. Med. Cell Longev. 2021, 2021, 6691226. [Google Scholar] [CrossRef] [PubMed]
  83. Crawford, P.A.; Schaffer, J.E. Metabolic stress in the myocardium: Adaptations of gene expression. J. Mol. Cell. Cardiol. 2013, 55, 130–138. [Google Scholar] [CrossRef] [PubMed]
  84. Romuk, E.; Jacheć, W.; Kozielska-Nowalany, E.; Birkner, E.; Zemła-Woszek, A.; Wojciechowska, C. Superoxide dismutase activity as a predictor of adverse outcomes in patients with nonischemic dilated cardiomyopathy. Cell Stress Chaperones 2019, 24, 661–673. [Google Scholar] [CrossRef] [PubMed]
  85. Guzik, T.J.; Touyz, R.M. Oxidative Stress, Inflammation, and Vascular Aging in Hypertension. Hypertension 2017, 70, 660–667. [Google Scholar] [CrossRef]
  86. Schulz, E.; Gori, T.; Münzel, T. Oxidative stress and endothelial dysfunction in hypertension. Hypertens. Res. 2011, 34, 665–673. [Google Scholar] [CrossRef]
  87. Case, A.J.; Tian, J.; Zimmerman, M.C. Increased mitochondrial superoxide in the brain, but not periphery, sensitizes mice to angiotensin II-mediated hypertension. Redox. Biol. 2017, 11, 82–90. [Google Scholar] [CrossRef]
  88. Chan, S.H.; Tai, M.H.; Li, C.Y.; Chan, J.Y. Reduction in molecular synthesis or enzyme activity of superoxide dismutases and catalase contributes to oxidative stress and neurogenic hypertension in spontaneously hypertensive rats. Free Radic. Biol. Med. 2006, 40, 2028–2039. [Google Scholar] [CrossRef]
  89. Li, L.; Crockett, E.; Wang, D.H.; Galligan, J.J.; Fink, G.D.; Chen, A.F. Gene transfer of endothelial NO synthase and manganese superoxide dismutase on arterial vascular cell adhesion molecule-1 expression and superoxide production in deoxycorticosterone acetate-salt hypertension. Arter. Thromb. Vasc. Biol. 2002, 22, 249–255. [Google Scholar] [CrossRef]
  90. Afolayan, A.J.; Eis, A.; Teng, R.J.; Bakhutashvili, I.; Kaul, S.; Davis, J.M.; Konduri, G.G. Decreases in manganese superoxide dismutase expression and activity contribute to oxidative stress in persistent pulmonary hypertension of the newborn. Am. J. Physiol. Lung Cell Mol. Physiol. 2012, 303, L870–L879. [Google Scholar] [CrossRef]
  91. Schultz, A.; Olorundami, O.A.; Teng, R.J.; Jarzembowski, J.; Shi, Z.Z.; Kumar, S.N.; Pritchard, K., Jr.; Konduri, G.G.; Afolayan, A.J. Decreased OLA1 (Obg-Like ATPase-1) Expression Drives Ubiquitin-Proteasome Pathways to Downregulate Mitochondrial SOD2 (Superoxide Dismutase) in Persistent Pulmonary Hypertension of the Newborn. Hypertension 2019, 74, 957–966. [Google Scholar] [CrossRef] [PubMed]
  92. Yeh, H.L.; Kuo, L.T.; Sung, F.C.; Yeh, C.C. Association between Polymorphisms of Antioxidant Gene (MnSOD, CAT, and GPx1) and Risk of Coronary Artery Disease. Biomed. Res. Int. 2018, 2018, 5086869. [Google Scholar] [CrossRef] [PubMed]
  93. Souiden, Y.; Mallouli, H.; Meskhi, S.; Chaabouni, Y.; Rebai, A.; Chéour, F.; Mahdouani, K. MnSOD and GPx1 polymorphism relationship with coronary heart disease risk and severity. Biol. Res. 2016, 49, 22. [Google Scholar] [CrossRef]
  94. Peoples, J.N.; Saraf, A.; Ghazal, N.; Pham, T.T.; Kwong, J.Q. Mitochondrial dysfunction and oxidative stress in heart disease. Exp. Mol. Med. 2019, 51, 1–13. [Google Scholar] [CrossRef]
  95. Loch, T.; Vakhrusheva, O.; Piotrowska, I.; Ziolkowski, W.; Ebelt, H.; Braun, T.; Bober, E. Different extent of cardiac malfunction and resistance to oxidative stress in heterozygous and homozygous manganese-dependent superoxide dismutase-mutant mice. Cardiovasc. Res. 2009, 82, 448–457. [Google Scholar] [CrossRef] [PubMed]
  96. Roos, C.M.; Zhang, B.; Hagler, M.A.; Verzosa, G.C.; Huang, R.; Oehler, E.A.; Arghami, A.; Miller, J.D. Effects of Altering Mitochondrial Antioxidant Capacity on Molecular and Phenotypic Drivers of Fibrocalcific Aortic Valve Stenosis. Front. Cardiovasc. Med. 2021, 8, 694881. [Google Scholar] [CrossRef]
  97. Ohashi, M.; Runge, M.S.; Faraci, F.M.; Heistad, D.D. MnSOD deficiency increases endothelial dysfunction in ApoE-deficient mice. Arter. Thromb. Vasc. Biol. 2006, 26, 2331–2336. [Google Scholar] [CrossRef]
  98. Ballinger, S.W.; Patterson, C.; Knight-Lozano, C.A.; Burow, D.L.; Conklin, C.A.; Hu, Z.; Reuf, J.; Horaist, C.; Lebovitz, R.; Hunter, G.C.; et al. Mitochondrial integrity and function in atherogenesis. Circulation 2002, 106, 544–549. [Google Scholar] [CrossRef]
  99. Harrison, C.M.; Pompilius, M.; Pinkerton, K.E.; Ballinger, S.W. Mitochondrial oxidative stress significantly influences atherogenic risk and cytokine-induced oxidant production. Environ. Health Perspect. 2011, 119, 676–681. [Google Scholar] [CrossRef]
  100. Daosukho, C.; Ittarat, W.; Lin, S.-m.; Sawyer, D.B.; Kiningham, K.; Lien, Y.-C.; Clair, D.K.S. Induction of manganese superoxide dismutase (MnSOD) mediates cardioprotective effect of tamoxifen (TAM). J. Mol. Cell. Cardiol. 2005, 39, 792–803. [Google Scholar] [CrossRef]
  101. Jain, K.; Gu, S.X.; Hwa, J. SOD2 in platelets: With age comes responsibility. J. Thromb. Haemost. 2023, 21, 1077–1081. [Google Scholar] [CrossRef] [PubMed]
  102. Teixeira, R.B.; Pfeiffer, M.; Zhang, P.; Shafique, E.; Rayta, B.; Karbasiafshar, C.; Ahsan, N.; Sellke, F.W.; Abid, M.R. Reduction in mitochondrial ROS improves oxidative phosphorylation and provides resilience to coronary endothelium in non-reperfused myocardial infarction. Basic Res. Cardiol. 2023, 118, 3. [Google Scholar] [CrossRef] [PubMed]
  103. Liu, H.; Wang, Q.; Liu, D.; Li, Z.; Fu, Y.; Tse, G.; Li, G.; Liu, T.; Xu, G. Manganese Superoxide Dismutase as a Novel Oxidative Stress Biomarker for Predicting Paroxysmal Atrial Fibrillation. J. Clin. Med. 2022, 11, 5131. [Google Scholar] [CrossRef]
  104. Gong, J.; Sun, P.; Li, L.; Zou, Z.; Wu, Q.; Sun, L.; Li, H.; Gu, Z.; Su, L. Heat stress suppresses MnSOD expression via p53-Sp1 interaction and induces oxidative stress damage in endothelial cells: Protective effects of MitoQ10 and Pifithrin-α. Heliyon 2023, 9, e22805. [Google Scholar] [CrossRef]
  105. Zou, X.; Zhu, Y.; Park, S.H.; Liu, G.; O’Brien, J.; Jiang, H.; Gius, D. SIRT3-Mediated Dimerization of IDH2 Directs Cancer Cell Metabolism and Tumor Growth. Cancer Res. 2017, 77, 3990–3999. [Google Scholar] [CrossRef] [PubMed]
  106. Pollard, P.J.; Brière, J.J.; Alam, N.A.; Barwell, J.; Barclay, E.; Wortham, N.C.; Hunt, T.; Mitchell, M.; Olpin, S.; Moat, S.J.; et al. Accumulation of Krebs cycle intermediates and over-expression of HIF1alpha in tumours which result from germline FH and SDH mutations. Hum. Mol. Genet. 2005, 14, 2231–2239. [Google Scholar] [CrossRef]
  107. Hao, H.X.; Khalimonchuk, O.; Schraders, M.; Dephoure, N.; Bayley, J.P.; Kunst, H.; Devilee, P.; Cremers, C.W.; Schiffman, J.D.; Bentz, B.G.; et al. SDH5, a gene required for flavination of succinate dehydrogenase, is mutated in paraganglioma. Science 2009, 325, 1139–1142. [Google Scholar] [CrossRef]
  108. Selak, M.A.; Armour, S.M.; MacKenzie, E.D.; Boulahbel, H.; Watson, D.G.; Mansfield, K.D.; Pan, Y.; Simon, M.C.; Thompson, C.B.; Gottlieb, E. Succinate links TCA cycle dysfunction to oncogenesis by inhibiting HIF-alpha prolyl hydroxylase. Cancer Cell 2005, 7, 77–85. [Google Scholar] [CrossRef]
  109. Palma, F.R.; He, C.; Danes, J.M.; Paviani, V.; Coelho, D.R.; Gantner, B.N.; Bonini, M.G. Mitochondrial Superoxide Dismutase: What the Established, the Intriguing, and the Novel Reveal About a Key Cellular Redox Switch. Antioxid. Redox. Signal 2020, 32, 701–714. [Google Scholar] [CrossRef]
  110. Dhar, S.K.; Tangpong, J.; Chaiswing, L.; Oberley, T.D.; St Clair, D.K. Manganese superoxide dismutase is a p53-regulated gene that switches cancers between early and advanced stages. Cancer Res. 2011, 71, 6684–6695. [Google Scholar] [CrossRef]
  111. Zhao, Y.; Oberley, T.D.; Chaiswing, L.; Lin, S.M.; Epstein, C.J.; Huang, T.T.; St Clair, D. Manganese superoxide dismutase deficiency enhances cell turnover via tumor promoter-induced alterations in AP-1 and p53-mediated pathways in a skin cancer model. Oncogene 2002, 21, 3836–3846. [Google Scholar] [CrossRef] [PubMed]
  112. Darby Weydert, C.J.; Smith, B.B.; Xu, L.; Kregel, K.C.; Ritchie, J.M.; Davis, C.S.; Oberley, L.W. Inhibition of oral cancer cell growth by adenovirusMnSOD plus BCNU treatment. Free Radic. Biol. Med. 2003, 34, 316–329. [Google Scholar] [CrossRef] [PubMed]
  113. Ough, M.; Lewis, A.; Zhang, Y.; Hinkhouse, M.M.; Ritchie, J.M.; Oberley, L.W.; Cullen, J.J. Inhibition of cell growth by overexpression of manganese superoxide dismutase (MnSOD) in human pancreatic carcinoma. Free Radic. Res. 2004, 38, 1223–1233. [Google Scholar] [CrossRef]
  114. Zhong, W.; Oberley, L.W.; Oberley, T.D.; Yan, T.; Domann, F.E.; St Clair, D.K. Inhibition of cell growth and sensitization to oxidative damage by overexpression of manganese superoxide dismutase in rat glioma cells. Cell Growth Differ. 1996, 7, 1175–1186. [Google Scholar]
  115. Zhong, W.; Yan, T.; Webber, M.M.; Oberley, T.D. Alteration of cellular phenotype and responses to oxidative stress by manganese superoxide dismutase and a superoxide dismutase mimic in RWPE-2 human prostate adenocarcinoma cells. Antioxid. Redox. Signal 2004, 6, 513–522. [Google Scholar] [CrossRef]
  116. Paku, M.; Haraguchi, N.; Takeda, M.; Fujino, S.; Ogino, T.; Takahashi, H.; Miyoshi, N.; Uemura, M.; Mizushima, T.; Yamamoto, H.; et al. SIRT3-Mediated SOD2 and PGC-1α Contribute to Chemoresistance in Colorectal Cancer Cells. Ann. Surg. Oncol. 2021, 28, 4720–4732. [Google Scholar] [CrossRef]
  117. Hsueh, W.T.; Chen, S.H.; Chien, C.H.; Chou, S.W.; Chi, P.I.; Chu, J.M.; Chang, K.Y. SOD2 Enhancement by Long-Term Inhibition of the PI3K Pathway Confers Multi-Drug Resistance and Enhanced Tumor-Initiating Features in Head and Neck Cancer. Int. J. Mol. Sci. 2021, 22, 11260. [Google Scholar] [CrossRef] [PubMed]
  118. Zhang, Y.; Xu, Z.; Ding, J.; Tan, C.; Hu, W.; Li, Y.; Huang, W.; Xu, Y. HZ08 suppresses RelB-activated MnSOD expression and enhances Radiosensitivity of prostate Cancer cells. J. Exp. Clin. Cancer Res. 2018, 37, 174. [Google Scholar] [CrossRef]
  119. Noh, J.K.; Woo, S.R.; Yun, M.; Lee, M.K.; Kong, M.; Min, S.; Kim, S.I.; Lee, Y.C.; Eun, Y.G.; Ko, S.G. SOD2- and NRF2-associated Gene Signature to Predict Radioresistance in Head and Neck Cancer. Cancer Genom. Proteom. 2021, 18, 675–684. [Google Scholar] [CrossRef]
  120. Ma, C.S.; Lv, Q.M.; Zhang, K.R.; Tang, Y.B.; Zhang, Y.F.; Shen, Y.; Lei, H.M.; Zhu, L. NRF2-GPX4/SOD2 axis imparts resistance to EGFR-tyrosine kinase inhibitors in non-small-cell lung cancer cells. Acta Pharmacol. Sin. 2021, 42, 613–623. [Google Scholar] [CrossRef]
  121. Ollila, E.; Ahtikoski, A.; Puistola, U.; Karihtala, P.; Urpilainen, E. Redox-state-regulating enzymes have prognostic value in diabetic endometrial cancer patients: Impact of statin use? Front. Oncol. 2024, 14, 1393103. [Google Scholar] [CrossRef]
  122. Piecuch, J.Z.; Kucharzewski, M.; Wyrobiec, G.; Brzozowa-Zasada, M. Immunohistochemical detection of MnSOD in colon adenocarcinoma patients—Clinical application. Prz. Gastroenterol. 2024, 19, 186–193. [Google Scholar] [CrossRef] [PubMed]
  123. Nie, S.; Shi, Z.; Shi, M.; Li, H.; Qian, X.; Peng, C.; Ding, X.; Zhang, S.; Lv, Y.; Wang, L.; et al. PPARγ/SOD2 Protects Against Mitochondrial ROS-Dependent Apoptosis via Inhibiting ATG4D-Mediated Mitophagy to Promote Pancreatic Cancer Proliferation. Front. Cell Dev. Biol. 2021, 9, 745554. [Google Scholar] [CrossRef]
  124. Kim, Y.S.; Gupta Vallur, P.; Phaëton, R.; Mythreye, K.; Hempel, N. Insights into the Dichotomous Regulation of SOD2 in Cancer. Antioxidants 2017, 6, 86. [Google Scholar] [CrossRef] [PubMed]
  125. Walker, G.J.; Palmer, J.M.; Walters, M.K.; Nancarrow, D.J.; Parsons, P.G.; Hayward, N.K. Simple tandem repeat allelic deletions confirm the preferential loss of distal chromosome 6q in melanoma. Int. J. Cancer 1994, 58, 203–206. [Google Scholar] [CrossRef]
  126. Bravard, A.; Hoffschir, F.; Sabatier, L.; Ricoul, M.; Pinton, A.; Cassingena, R.; Estrade, S.; Luccioni, C.; Dutrillaux, B. Early superoxide dismutase alterations during SV40-transformation of human fibroblasts. Int. J. Cancer 1992, 52, 797–801. [Google Scholar] [CrossRef] [PubMed]
  127. Xu, Y.; Fang, F.; Dhar, S.K.; Bosch, A.; St Clair, W.H.; Kasarskis, E.J.; St Clair, D.K. Mutations in the SOD2 promoter reveal a molecular basis for an activating protein 2-dependent dysregulation of manganese superoxide dismutase expression in cancer cells. Mol. Cancer Res. 2008, 6, 1881–1893. [Google Scholar] [CrossRef]
  128. Douglas, D.B.; Akiyama, Y.; Carraway, H.; Belinsky, S.A.; Esteller, M.; Gabrielson, E.; Weitzman, S.; Williams, T.; Herman, J.G.; Baylin, S.B. Hypermethylation of a small CpGuanine-rich region correlates with loss of activator protein-2alpha expression during progression of breast cancer. Cancer Res. 2004, 64, 1611–1620. [Google Scholar] [CrossRef]
  129. Xu, Y.; Krishnan, A.; Wan, X.S.; Majima, H.; Yeh, C.C.; Ludewig, G.; Kasarskis, E.J.; St Clair, D.K. Mutations in the promoter reveal a cause for the reduced expression of the human manganese superoxide dismutase gene in cancer cells. Oncogene 1999, 18, 93–102. [Google Scholar] [CrossRef]
  130. Minig, V.; Kattan, Z.; van Beeumen, J.; Brunner, E.; Becuwe, P. Identification of DDB2 protein as a transcriptional regulator of constitutive SOD2 gene expression in human breast cancer cells. J. Biol. Chem. 2009, 284, 14165–14176. [Google Scholar] [CrossRef]
  131. Hitchler, M.J.; Wikainapakul, K.; Yu, L.; Powers, K.; Attatippaholkun, W.; Domann, F.E. Epigenetic regulation of manganese superoxide dismutase expression in human breast cancer cells. Epigenetics 2006, 1, 163–171. [Google Scholar] [CrossRef] [PubMed]
  132. Hodge, D.R.; Peng, B.; Pompeia, C.; Thomas, S.; Cho, E.; Clausen, P.A.; Marquez, V.E.; Farrar, W.L. Epigenetic silencing of manganese superoxide dismutase (SOD-2) in KAS 6/1 human multiple myeloma cells increases cell proliferation. Cancer Biol. Ther. 2005, 4, 585–592. [Google Scholar] [CrossRef]
  133. Hurt, E.M.; Thomas, S.B.; Peng, B.; Farrar, W.L. Integrated molecular profiling of SOD2 expression in multiple myeloma. Blood 2007, 109, 3953–3962. [Google Scholar] [CrossRef]
  134. Hurt, E.M.; Thomas, S.B.; Peng, B.; Farrar, W.L. Molecular consequences of SOD2 expression in epigenetically silenced pancreatic carcinoma cell lines. Br. J. Cancer 2007, 97, 1116–1123. [Google Scholar] [CrossRef]
  135. Rivlin, N.; Brosh, R.; Oren, M.; Rotter, V. Mutations in the p53 Tumor Suppressor Gene: Important Milestones at the Various Steps of Tumorigenesis. Genes Cancer 2011, 2, 466–474. [Google Scholar] [CrossRef] [PubMed]
  136. Drane, P.; Bravard, A.; Bouvard, V.; May, E. Reciprocal down-regulation of p53 and SOD2 gene expression-implication in p53 mediated apoptosis. Oncogene 2001, 20, 430–439. [Google Scholar] [CrossRef]
  137. Pani, G.; Bedogni, B.; Anzevino, R.; Colavitti, R.; Palazzotti, B.; Borrello, S.; Galeotti, T. Deregulated manganese superoxide dismutase expression and resistance to oxidative injury in p53-deficient cells. Cancer Res. 2000, 60, 4654–4660. [Google Scholar] [PubMed]
  138. Hussain, S.P.; Amstad, P.; He, P.; Robles, A.; Lupold, S.; Kaneko, I.; Ichimiya, M.; Sengupta, S.; Mechanic, L.; Okamura, S.; et al. p53-induced up-regulation of MnSOD and GPx but not catalase increases oxidative stress and apoptosis. Cancer Res. 2004, 64, 2350–2356. [Google Scholar] [CrossRef]
  139. Dhar, S.K.; Xu, Y.; Chen, Y.; St Clair, D.K. Specificity protein 1-dependent p53-mediated suppression of human manganese superoxide dismutase gene expression. J. Biol. Chem. 2006, 281, 21698–21709. [Google Scholar] [CrossRef]
  140. Zhao, Y.; Chaiswing, L.; Velez, J.M.; Batinic-Haberle, I.; Colburn, N.H.; Oberley, T.D.; St Clair, D.K. p53 translocation to mitochondria precedes its nuclear translocation and targets mitochondrial oxidative defense protein-manganese superoxide dismutase. Cancer Res. 2005, 65, 3745–3750. [Google Scholar] [CrossRef]
  141. Li, H.; Wang, H.; Li, Z.; Kelley, N.; Ouyang, M.; Wu, J.W.; Meng, F.; Ou, W.B. Anti-proliferative and anti-invasive effects of exogenous thermostable MnSOD in gastric cancer associated with p53 and ZEB1 expression. J. Cancer 2025, 16, 2062–2074. [Google Scholar] [CrossRef]
  142. Xu, Y.; Kiningham, K.K.; Devalaraja, M.N.; Yeh, C.C.; Majima, H.; Kasarskis, E.J.; St Clair, D.K. An intronic NF-kappaB element is essential for induction of the human manganese superoxide dismutase gene by tumor necrosis factor-alpha and interleukin-1beta. DNA Cell Biol. 1999, 18, 709–722. [Google Scholar] [CrossRef] [PubMed]
  143. Sun, Y.; St Clair, D.K.; Fang, F.; Warren, G.W.; Rangnekar, V.M.; Crooks, P.A.; St Clair, W.H. The radiosensitization effect of parthenolide in prostate cancer cells is mediated by nuclear factor-kappaB inhibition and enhanced by the presence of PTEN. Mol. Cancer Ther. 2007, 6, 2477–2486. [Google Scholar] [CrossRef] [PubMed]
  144. Kamarajugadda, S.; Cai, Q.; Chen, H.; Nayak, S.; Zhu, J.; He, M.; Jin, Y.; Zhang, Y.; Ai, L.; Martin, S.S.; et al. Manganese superoxide dismutase promotes anoikis resistance and tumor metastasis. Cell Death Dis. 2013, 4, e504. [Google Scholar] [CrossRef] [PubMed]
  145. Delhalle, S.; Deregowski, V.; Benoit, V.; Merville, M.P.; Bours, V. NF-kappaB-dependent MnSOD expression protects adenocarcinoma cells from TNF-alpha-induced apoptosis. Oncogene 2002, 21, 3917–3924. [Google Scholar] [CrossRef]
  146. Nguyen, T.; Sherratt, P.J.; Pickett, C.B. Regulatory mechanisms controlling gene expression mediated by the antioxidant response element. Annu. Rev. Pharmacol. Toxicol. 2003, 43, 233–260. [Google Scholar] [CrossRef]
  147. Venugopal, R.; Jaiswal, A.K. Nrf1 and Nrf2 positively and c-Fos and Fra1 negatively regulate the human antioxidant response element-mediated expression of NAD(P)H:quinone oxidoreductase1 gene. Proc. Natl. Acad. Sci. USA 1996, 93, 14960–14965. [Google Scholar] [CrossRef]
  148. DeNicola, G.M.; Karreth, F.A.; Humpton, T.J.; Gopinathan, A.; Wei, C.; Frese, K.; Mangal, D.; Yu, K.H.; Yeo, C.J.; Calhoun, E.S.; et al. Oncogene-induced Nrf2 transcription promotes ROS detoxification and tumorigenesis. Nature 2011, 475, 106–109. [Google Scholar] [CrossRef]
  149. Kobayashi, A.; Kang, M.I.; Watai, Y.; Tong, K.I.; Shibata, T.; Uchida, K.; Yamamoto, M. Oxidative and electrophilic stresses activate Nrf2 through inhibition of ubiquitination activity of Keap1. Mol. Cell. Biol. 2006, 26, 221–229. [Google Scholar] [CrossRef]
  150. Dinkova-Kostova, A.T.; Holtzclaw, W.D.; Cole, R.N.; Itoh, K.; Wakabayashi, N.; Katoh, Y.; Yamamoto, M.; Talalay, P. Direct evidence that sulfhydryl groups of Keap1 are the sensors regulating induction of phase 2 enzymes that protect against carcinogens and oxidants. Proc. Natl. Acad. Sci. USA 2002, 99, 11908–11913. [Google Scholar] [CrossRef]
  151. Konstantinopoulos, P.A.; Spentzos, D.; Fountzilas, E.; Francoeur, N.; Sanisetty, S.; Grammatikos, A.P.; Hecht, J.L.; Cannistra, S.A. Keap1 mutations and Nrf2 pathway activation in epithelial ovarian cancer. Cancer Res. 2011, 71, 5081–5089. [Google Scholar] [CrossRef] [PubMed]
  152. Taylor, R.C.; Acquaah-Mensah, G.; Singhal, M.; Malhotra, D.; Biswal, S. Network inference algorithms elucidate Nrf2 regulation of mouse lung oxidative stress. PLoS Comput. Biol. 2008, 4, e1000166. [Google Scholar] [CrossRef]
  153. Solis, L.M.; Behrens, C.; Dong, W.; Suraokar, M.; Ozburn, N.C.; Moran, C.A.; Corvalan, A.H.; Biswal, S.; Swisher, S.G.; Bekele, B.N.; et al. Nrf2 and Keap1 abnormalities in non-small cell lung carcinoma and association with clinicopathologic features. Clin. Cancer Res. 2010, 16, 3743–3753. [Google Scholar] [CrossRef] [PubMed]
  154. Hanada, N.; Takahata, T.; Zhou, Q.; Ye, X.; Sun, R.; Itoh, J.; Ishiguro, A.; Kijima, H.; Mimura, J.; Itoh, K.; et al. Methylation of the KEAP1 gene promoter region in human colorectal cancer. BMC Cancer 2012, 12, 66. [Google Scholar] [CrossRef] [PubMed]
  155. Kops, G.J.; Dansen, T.B.; Polderman, P.E.; Saarloos, I.; Wirtz, K.W.; Coffer, P.J.; Huang, T.T.; Bos, J.L.; Medema, R.H.; Burgering, B.M. Forkhead transcription factor FOXO3a protects quiescent cells from oxidative stress. Nature 2002, 419, 316–321. [Google Scholar] [CrossRef]
  156. Li, M.; Chiu, J.F.; Mossman, B.T.; Fukagawa, N.K. Down-regulation of manganese-superoxide dismutase through phosphorylation of FOXO3a by Akt in explanted vascular smooth muscle cells from old rats. J. Biol. Chem. 2006, 281, 40429–40439. [Google Scholar] [CrossRef]
  157. Kato, M.; Yuan, H.; Xu, Z.G.; Lanting, L.; Li, S.L.; Wang, M.; Hu, M.C.; Reddy, M.A.; Natarajan, R. Role of the Akt/FoxO3a pathway in TGF-beta1-mediated mesangial cell dysfunction: A novel mechanism related to diabetic kidney disease. J. Am. Soc. Nephrol. 2006, 17, 3325–3335. [Google Scholar] [CrossRef]
  158. Jacobs, K.M.; Pennington, J.D.; Bisht, K.S.; Aykin-Burns, N.; Kim, H.S.; Mishra, M.; Sun, L.; Nguyen, P.; Ahn, B.H.; Leclerc, J.; et al. SIRT3 interacts with the daf-16 homolog FOXO3a in the mitochondria, as well as increases FOXO3a dependent gene expression. Int. J. Biol. Sci. 2008, 4, 291–299. [Google Scholar] [CrossRef]
  159. Lim, S.W.; Chen, W.C.; Ko, H.J.; Su, Y.F.; Wu, C.H.; Huang, F.L.; Li, C.F.; Tsai, C.Y. 6-Gingerol Induced Apoptosis and Cell Cycle Arrest in Glioma Cells via MnSOD and ERK Phosphorylation Modulation. Biomol. Ther. 2025, 33, 129–142. [Google Scholar] [CrossRef]
  160. Basilotta, R.; Lanza, M.; Filippone, A.; Casili, G.; Mannino, D.; De Gaetano, F.; Chisari, G.; Colarossi, L.; Motta, G.; Campolo, M.; et al. Therapeutic Potential of Dimethyl Fumarate in Counteract Oral Squamous Cell Carcinoma Progression by Modulating Apoptosis, Oxidative Stress and Epithelial-Mesenchymal Transition. Int. J. Mol. Sci. 2023, 24, 2777. [Google Scholar] [CrossRef]
  161. Kang, B.G.; Shende, M.; Inci, G.; Park, S.H.; Jung, J.S.; Kim, S.B.; Kim, J.H.; Mo, Y.W.; Seo, J.H.; Feng, J.H.; et al. Combination of metformin/efavirenz/fluoxetine exhibits profound anticancer activity via a cancer cell-specific ROS amplification. Cancer Biol. Ther. 2023, 24, 20–32. [Google Scholar] [CrossRef]
  162. Gao, Y.H.; Li, C.X.; Shen, S.M.; Li, H.; Chen, G.Q.; Wei, Q.; Wang, L.S. Hypoxia-inducible factor 1α mediates the down-regulation of superoxide dismutase 2 in von Hippel-Lindau deficient renal clear cell carcinoma. Biochem. Biophys. Res. Commun. 2013, 435, 46–51. [Google Scholar] [CrossRef] [PubMed]
  163. Gu, C.; Gonzalez, J.; Zhang, T.; Kamel-Reid, S.; Wells, R.A. The aryl hydrocarbon receptor nuclear translocator (ARNT) modulates the antioxidant response in AML cells. Leuk Res. 2013, 37, 1750–1756. [Google Scholar] [CrossRef]
  164. Solari, C.; Vázquez Echegaray, C.; Cosentino, M.S.; Petrone, M.V.; Waisman, A.; Luzzani, C.; Francia, M.; Villodre, E.; Lenz, G.; Miriuka, S.; et al. Manganese Superoxide Dismutase Gene Expression Is Induced by Nanog and Oct4, Essential Pluripotent Stem Cells’ Transcription Factors. PLoS ONE 2015, 10, e0144336. [Google Scholar] [CrossRef] [PubMed]
  165. Iglesias-Bartolome, R.; Patel, V.; Cotrim, A.; Leelahavanichkul, K.; Molinolo, A.A.; Mitchell, J.B.; Gutkind, J.S. mTOR inhibition prevents epithelial stem cell senescence and protects from radiation-induced mucositis. Cell Stem. Cell 2012, 11, 401–414. [Google Scholar] [CrossRef] [PubMed]
  166. Epperly, M.W.; Bray, J.A.; Esocobar, P.; Bigbee, W.L.; Watkins, S.; Greenberger, J.S. Overexpression of the human manganese superoxide dismutase (MnSOD) transgene in subclones of murine hematopoietic progenitor cell line 32D cl 3 decreases irradiation-induced apoptosis but does not alter G2/M or G1/S phase cell cycle arrest. Radiat. Oncol. Investig. 1999, 7, 331–342. [Google Scholar] [CrossRef]
  167. Deng, X.; Ewton, D.Z.; Friedman, E. Mirk/Dyrk1B maintains the viability of quiescent pancreatic cancer cells by reducing levels of reactive oxygen species. Cancer Res. 2009, 69, 3317–3324. [Google Scholar] [CrossRef]
  168. Stuart, J.J.; Egry, L.A.; Wong, G.H.; Kaspar, R.L. The 3′ UTR of human MnSOD mRNA hybridizes to a small cytoplasmic RNA and inhibits gene expression. Biochem. Biophys. Res. Commun. 2000, 274, 641–648. [Google Scholar] [CrossRef]
  169. Pandit, H.; Zhang, W.; Li, Y.; Agle, S.; Li, X.; Li, S.P.; Cui, G.; Li, Y.; Martin, R.C. Manganese superoxide dismutase expression is negatively associated with microRNA-301a in human pancreatic ductal adenocarcinoma. Cancer Gene Ther. 2015, 22, 481–486. [Google Scholar] [CrossRef]
  170. Xu, Y.; Fang, F.; Zhang, J.; Josson, S.; St Clair, W.H.; St Clair, D.K. miR-17* suppresses tumorigenicity of prostate cancer by inhibiting mitochondrial antioxidant enzymes. PLoS ONE 2010, 5, e14356. [Google Scholar] [CrossRef]
  171. Liu, X.; Yu, J.; Jiang, L.; Wang, A.; Shi, F.; Ye, H.; Zhou, X. MicroRNA-222 regulates cell invasion by targeting matrix metalloproteinase 1 (MMP1) and manganese superoxide dismutase 2 (SOD2) in tongue squamous cell carcinoma cell lines. Cancer Genom. Proteom. 2009, 6, 131–139. [Google Scholar]
  172. Kriegel, A.J.; Fang, Y.; Liu, Y.; Tian, Z.; Mladinov, D.; Matus, I.R.; Ding, X.; Greene, A.S.; Liang, M. MicroRNA-target pairs in human renal epithelial cells treated with transforming growth factor beta 1: A novel role of miR-382. Nucleic Acids Res. 2010, 38, 8338–8347. [Google Scholar] [CrossRef]
  173. Cui, Y.; She, K.; Tian, D.; Zhang, P.; Xin, X. miR-146a Inhibits Proliferation and Enhances Chemosensitivity in Epithelial Ovarian Cancer via Reduction of SOD2. Oncol. Res. 2016, 23, 275–282. [Google Scholar] [CrossRef] [PubMed]
  174. Miao, L.; St Clair, D.K. Regulation of superoxide dismutase genes: Implications in disease. Free Radic. Biol. Med. 2009, 47, 344–356. [Google Scholar] [CrossRef]
  175. Hernandez-Saavedra, D.; McCord, J.M. Paradoxical effects of thiol reagents on Jurkat cells and a new thiol-sensitive mutant form of human mitochondrial superoxide dismutase. Cancer Res. 2003, 63, 159–163. [Google Scholar]
  176. Zhang, H.J.; Yan, T.; Oberley, T.D.; Oberley, L.W. Comparison of effects of two polymorphic variants of manganese superoxide dismutase on human breast MCF-7 cancer cell phenotype. Cancer Res. 1999, 59, 6276–6283. [Google Scholar] [PubMed]
  177. Liu, G.; Zhou, W.; Park, S.; Wang, L.I.; Miller, D.P.; Wain, J.C.; Lynch, T.J.; Su, L.; Christiani, D.C. The SOD2 Val/Val genotype enhances the risk of nonsmall cell lung carcinoma by p53 and XRCC1 polymorphisms. Cancer 2004, 101, 2802–2808. [Google Scholar] [CrossRef]
  178. Wheatley-Price, P.; Asomaning, K.; Reid, A.; Zhai, R.; Su, L.; Zhou, W.; Zhu, A.; Ryan, D.P.; Christiani, D.C.; Liu, G. Myeloperoxidase and superoxide dismutase polymorphisms are associated with an increased risk of developing pancreatic adenocarcinoma. Cancer 2008, 112, 1037–1042. [Google Scholar] [CrossRef]
  179. Becuwe, P.; Ennen, M.; Klotz, R.; Barbieux, C.; Grandemange, S. Manganese superoxide dismutase in breast cancer: From molecular mechanisms of gene regulation to biological and clinical significance. Free Radic. Biol. Med. 2014, 77, 139–151. [Google Scholar] [CrossRef]
  180. Bag, A.; Bag, N. Target sequence polymorphism of human manganese superoxide dismutase gene and its association with cancer risk: A review. Cancer Epidemiol. Biomark. Prev. 2008, 17, 3298–3305. [Google Scholar] [CrossRef]
  181. Moradabadi, A.; Fekri-Soofiabadi, M.; Soltani, A.; Dabiri, S. An investigation into the manganese superoxide dismutase (MnSOD Val-9Ala) gene polymorphisms employing high-resolution melting in patients with gastric cancer: A preliminary study. Cancer Treat. Res. Commun. 2025, 44, 100942. [Google Scholar] [CrossRef]
  182. Zhu, Y.; Zou, X.; Dean, A.E.; Brien, J.O.; Gao, Y.; Tran, E.L.; Park, S.H.; Liu, G.; Kieffer, M.B.; Jiang, H.; et al. Lysine 68 acetylation directs MnSOD as a tetrameric detoxification complex versus a monomeric tumor promoter. Nat. Commun. 2019, 10, 2399. [Google Scholar] [CrossRef] [PubMed]
  183. He, C.; Danes, J.M.; Hart, P.C.; Zhu, Y.; Huang, Y.; de Abreu, A.L.; O’Brien, J.; Mathison, A.J.; Tang, B.; Frasor, J.M.; et al. SOD2 acetylation on lysine 68 promotes stem cell reprogramming in breast cancer. Proc. Natl. Acad. Sci. USA 2019, 116, 23534–23541. [Google Scholar] [CrossRef]
  184. Dhar, S.K.; St Clair, D.K. Manganese superoxide dismutase regulation and cancer. Free Radic. Biol. Med. 2012, 52, 2209–2222. [Google Scholar] [CrossRef]
  185. Gatenby, R.A.; Gillies, R.J. Why do cancers have high aerobic glycolysis? Nat. Rev. Cancer 2004, 4, 891–899. [Google Scholar] [CrossRef] [PubMed]
  186. Hanahan, D. Hallmarks of Cancer: New Dimensions. Cancer Discov. 2022, 12, 31–46. [Google Scholar] [CrossRef]
  187. Faubert, B.; Solmonson, A.; DeBerardinis, R.J. Metabolic reprogramming and cancer progression. Science 2020, 368, eaaw5473. [Google Scholar] [CrossRef] [PubMed]
  188. Zhou, C.; Lyu, L.H.; Miao, H.K.; Bahr, T.; Zhang, Q.Y.; Liang, T.; Zhou, H.B.; Chen, G.R.; Bai, Y. Redox regulation by SOD2 modulates colorectal cancer tumorigenesis through AMPK-mediated energy metabolism. Mol. Carcinog. 2020, 59, 545–556. [Google Scholar] [CrossRef]
  189. Hart, P.C.; Mao, M.; de Abreu, A.L.; Ansenberger-Fricano, K.; Ekoue, D.N.; Ganini, D.; Kajdacsy-Balla, A.; Diamond, A.M.; Minshall, R.D.; Consolaro, M.E.; et al. MnSOD upregulation sustains the Warburg effect via mitochondrial ROS and AMPK-dependent signalling in cancer. Nat. Commun. 2015, 6, 6053. [Google Scholar] [CrossRef]
  190. Quiros-Gonzalez, I.; Gonzalez-Menendez, P.; Mayo, J.C.; Hevia, D.; Artime-Naveda, F.; Fernandez-Vega, S.; Fernandez-Fernandez, M.; Rodriguez-Gonzalez, P.; Garcia-Alonso, J.I.; Sainz, R.M. Androgen-Dependent Prostate Cancer Cells Reprogram Their Metabolic Signature upon GLUT1 Upregulation by Manganese Superoxide Dismutase. Antioxidants 2022, 11, 313. [Google Scholar] [CrossRef]
  191. Liu, Y.D.; Yu, L.; Ying, L.; Balic, J.; Gao, H.; Deng, N.T.; West, A.; Yan, F.; Ji, C.B.; Gough, D.; et al. Toll-like receptor 2 regulates metabolic reprogramming in gastric cancer via superoxide dismutase 2. Int. J. Cancer 2019, 144, 3056–3069. [Google Scholar] [CrossRef] [PubMed]
  192. Ibrahim, N.K.; Schreek, S.; Cinar, B.; Stasche, A.S.; Lee, S.H.; Zeug, A.; Dolgner, T.; Niessen, J.; Ponimaskin, E.; Shcherbata, H.; et al. SOD2 is a regulator of proteasomal degradation promoting an adaptive cellular starvation response. Cell Rep. 2025, 44, 115434. [Google Scholar] [CrossRef] [PubMed]
  193. Su, H.; Cai, T.; Zhang, S.; Yan, X.; Zhou, L.; He, Z.; Xue, P.; Li, J.; Zheng, M.; Yang, X.; et al. Identification of hub genes associated with neutrophils infiltration in colorectal cancer. J. Cell Mol. Med. 2021, 25, 3371–3380. [Google Scholar] [CrossRef] [PubMed]
  194. Han, X.; Liu, X.; Wang, X.; Guo, W.; Wen, Y.; Meng, W.; Peng, D.; Lv, P.; Zhang, X.; Shen, H. TNF-α-dependent lung inflammation upregulates superoxide dismutase-2 to promote tumor cell proliferation in lung adenocarcinoma. Mol. Carcinog. 2020, 59, 1088–1099. [Google Scholar] [CrossRef]
  195. Al Haq, A.T.; Tseng, H.Y.; Chen, L.M.; Wang, C.C.; Hsu, H.L. Targeting prooxidant MnSOD effect inhibits triple-negative breast cancer (TNBC) progression and M2 macrophage functions under the oncogenic stress. Cell Death Dis. 2022, 13, 49. [Google Scholar] [CrossRef]
  196. Lou, J.; Dong, J.; Xu, R.; Zeng, H.; Fang, L.; Wu, Y.; Liu, Y.; Wang, S. Remodeling of the tumor microenvironment using an engineered oncolytic vaccinia virus improves PD-L1 inhibition outcomes. Biosci. Rep. 2021, 41, BSR20204186. [Google Scholar] [CrossRef]
  197. Maier, C.M.; Chan, P.H. Role of superoxide dismutases in oxidative damage and neurodegenerative disorders. Neuroscientist 2002, 8, 323–334. [Google Scholar] [CrossRef]
  198. Pandey, M.K.; Mittra, P.; Maheshwari, P. The lipid peroxidation product as a marker of oxidative stress in epilepsy. J. Clin. Diagn. Res. 2012, 6, 590–592. [Google Scholar]
  199. Pearson, J.N.; Rowley, S.; Liang, L.P.; White, A.M.; Day, B.J.; Patel, M. Reactive oxygen species mediate cognitive deficits in experimental temporal lobe epilepsy. Neurobiol. Dis. 2015, 82, 289–297. [Google Scholar] [CrossRef]
  200. Kegler, A.; Caprara, A.L.F.; Pascotini, E.T.; Arend, J.; Gabbi, P.; Duarte, M.; Furian, A.F.; Oliveira, M.S.; Royes, L.F.F.; Fighera, M.R. Apoptotic Markers Are Increased in Epilepsy Patients: A Relation with Manganese Superoxide Dismutase Ala16Val Polymorphism and Seizure Type through IL-1β and IL-6 Pathways. Biomed. Res. Int. 2020, 2020, 6250429. [Google Scholar] [CrossRef]
  201. Andreassen, O.A.; Ferrante, R.J.; Dedeoglu, A.; Albers, D.W.; Klivenyi, P.; Carlson, E.J.; Epstein, C.J.; Beal, M.F. Mice with a partial deficiency of manganese superoxide dismutase show increased vulnerability to the mitochondrial toxins malonate, 3-nitropropionic acid, and MPTP. Exp. Neurol. 2001, 167, 189–195. [Google Scholar] [CrossRef] [PubMed]
  202. Suzuki, M.; Nagai, Y.; Wada, K.; Koike, T. Calcium leak through ryanodine receptor is involved in neuronal death induced by mutant huntingtin. Biochem. Biophys. Res. Commun. 2012, 429, 18–23. [Google Scholar] [CrossRef] [PubMed]
  203. Gonchar, O.O.; Maznychenko, A.V.; Klyuchko, O.M.; Mankovska, I.M.; Butowska, K.; Borowik, A.; Piosik, J.; Sokolowska, I. C(60) Fullerene Reduces 3-Nitropropionic Acid-Induced Oxidative Stress Disorders and Mitochondrial Dysfunction in Rats by Modulation of p53, Bcl-2 and Nrf2 Targeted Proteins. Int. J. Mol. Sci. 2021, 22, 5444. [Google Scholar] [CrossRef]
  204. Chen, H.; Song, Y.S.; Chan, P.H. Inhibition of NADPH oxidase is neuroprotective after ischemia-reperfusion. J. Cereb. Blood Flow Metab. 2009, 29, 1262–1272. [Google Scholar] [CrossRef] [PubMed]
  205. Corniola, R.; Zou, Y.; Leu, D.; Fike, J.R.; Huang, T.T. Paradoxical relationship between Mn superoxide dismutase deficiency and radiation-induced cognitive defects. PLoS ONE 2012, 7, e49367. [Google Scholar] [CrossRef]
  206. Jin, Z.; Zhang, Z.; Ke, J.; Wang, Y.; Wu, H. Exercise-Linked Irisin Prevents Mortality and Enhances Cognition in a Mice Model of Cerebral Ischemia by Regulating Klotho Expression. Oxid. Med. Cell Longev. 2021, 2021, 1697070. [Google Scholar] [CrossRef]
  207. Li, C.; Li, J.; Loreno, E.G.; Miriyala, S.; Panchatcharam, M.; Lu, X.; Sun, H. Chronic Low-Dose Alcohol Consumption Attenuates Post-Ischemic Inflammation via PPARγ in Mice. Int. J. Mol. Sci. 2021, 22, 5121. [Google Scholar] [CrossRef]
  208. Guo, S.; Li, F.; Wills, M.; Yip, J.; Wehbe, A.; Peng, C.; Geng, X.; Ding, Y. Chlorpromazine and Promethazine (C+P) Reduce Brain Injury after Ischemic Stroke through the PKC-δ/NOX/MnSOD Pathway. Mediat. Inflamm. 2022, 2022, 6886752. [Google Scholar] [CrossRef]
  209. Xie, W.; Zhu, T.; Zhou, P.; Xu, H.; Meng, X.; Ding, T.; Nan, F.; Sun, G.; Sun, X. Notoginseng leaf triterpenes ameliorates mitochondrial oxidative injury via the NAMPT-SIRT1/2/3 signaling pathways in cerebral ischemic model rats. J. Ginseng. Res. 2023, 47, 199–209. [Google Scholar] [CrossRef]
  210. Xie, W.; Zhu, T.; Zhou, P.; Xu, H.; Meng, X.; Ding, T.; Nan, F.; Sun, G.; Sun, X. Notoginseng Leaf Triterpenes Ameliorates OGD/R-Induced Neuronal Injury via SIRT1/2/3-Foxo3a-MnSOD/PGC-1α Signaling Pathways Mediated by the NAMPT-NAD Pathway. Oxid. Med. Cell Longev. 2020, 2020, 7308386. [Google Scholar] [CrossRef]
  211. Przedborski, S.; Kostic, V.; Jackson-Lewis, V.; Naini, A.B.; Simonetti, S.; Fahn, S.; Carlson, E.; Epstein, C.J.; Cadet, J.L. Transgenic mice with increased Cu/Zn-superoxide dismutase activity are resistant to N-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-induced neurotoxicity. J. Neurosci. 1992, 12, 1658–1667. [Google Scholar] [CrossRef]
  212. Filograna, R.; Godena, V.K.; Sanchez-Martinez, A.; Ferrari, E.; Casella, L.; Beltramini, M.; Bubacco, L.; Whitworth, A.J.; Bisaglia, M. Superoxide Dismutase (SOD)-mimetic M40403 Is Protective in Cell and Fly Models of Paraquat Toxicity: IMPLICATIONS FOR PARKINSON DISEASE. J. Biol. Chem. 2016, 291, 9257–9267. [Google Scholar] [CrossRef]
  213. Shi, H.; Deng, H.X.; Gius, D.; Schumacker, P.T.; Surmeier, D.J.; Ma, Y.C. Sirt3 protects dopaminergic neurons from mitochondrial oxidative stress. Hum. Mol. Genet. 2017, 26, 1915–1926. [Google Scholar] [CrossRef]
  214. Kim, D.Y.; Leem, Y.H.; Kim, H.S. MLKL Inhibitor Reduces Oxidative Stress, Inflammation, and Dopaminergic Neuronal Cell Death in MPTP-Induced Parkinson’s Disease Mouse Model. Biomol. Ther. 2025, 33, 429–437. [Google Scholar] [CrossRef]
  215. Massaad, C.A.; Washington, T.M.; Pautler, R.G.; Klann, E. Overexpression of SOD-2 reduces hippocampal superoxide and prevents memory deficits in a mouse model of Alzheimer’s disease. Proc. Natl. Acad. Sci. USA 2009, 106, 13576–13581. [Google Scholar] [CrossRef]
  216. Murakami, K.; Shimizu, T.; Irie, K. Formation of the 42-mer Amyloid β Radical and the Therapeutic Role of Superoxide Dismutase in Alzheimer’s Disease. J. Amino. Acids 2011, 2011, 654207. [Google Scholar] [CrossRef]
  217. Wang, H.; Zhang, Z.; Hongpaisan, J. PKCε activator protects hippocampal microvascular disruption and memory defect in 3×Tg-Alzheimer’s disease mice with cerebral microinfarcts. Front. Aging Neurosci. 2023, 15, 1272361. [Google Scholar] [CrossRef]
  218. Millien, G.; Wang, H.; Zhang, Z.; Alkon, D.L.; Hongpaisan, J. PKCε Activation Restores Loss of PKCε, Manganese Superoxide Dismutase, Vascular Endothelial Growth Factor, and Microvessels in Aged and Alzheimer’s Disease Hippocampus. Front. Aging Neurosci. 2022, 14, 836634. [Google Scholar] [CrossRef]
  219. Lanza, M.; Basilotta, R.; Cuzzocrea, S.; Bulzomì, M.; Oddo, S.; Casili, G.; Esposito, E. An RNAi-Mediated Reduction in Transcription Factor Nrf-2 Blocks the Positive Effects of Dimethyl Fumarate on Metabolic Stress in Alzheimer’s Disease. Int. J. Mol. Sci. 2023, 24, 11303. [Google Scholar] [CrossRef]
  220. Li, Y.; Lu, J.; Cao, X.; Zhao, H.; Gao, L.; Xia, P.; Pei, G. A Newly Synthesized Rhamnoside Derivative Alleviates Alzheimer’s Amyloid-β-Induced Oxidative Stress, Mitochondrial Dysfunction, and Cell Senescence through Upregulating SIRT3. Oxid. Med. Cell Longev. 2020, 2020, 7698560. [Google Scholar] [CrossRef]
  221. Patrick, S.; Corrigan, R.; Grizzanti, J.; Mey, M.; Blair, J.; Pallas, M.; Camins, A.; Lee, H.G.; Casadesus, G. Neuroprotective Effects of the Amylin Analog, Pramlintide, on Alzheimer’s Disease Are Associated with Oxidative Stress Regulation Mechanisms. J. Alzheimers Dis. 2019, 69, 157–168. [Google Scholar] [CrossRef]
  222. Wang, H.; Zhang, Z.; Sittirattanayeunyong, S.; Hongpaisan, J. Association of Apolipoprotein E4-related Microvascular Disease in the Alzheimer’s Disease Hippocampal CA1 Stratum Radiatum. Neuroscience 2023, 526, 204–222. [Google Scholar] [CrossRef]
  223. Villasana, L.E.; Akinyeke, T.; Weber, S.; Raber, J. Paradoxical effects of (137)Cs irradiation on pharmacological stimulation of reactive oxygen species in hippocampal slices from apoE2 and apoE4 mice. Oncotarget 2017, 8, 76587–76605. [Google Scholar] [CrossRef]
  224. Campagnoli, L.I.M.; Ahmad, L.; Marchesi, N.; Greco, G.; Boschi, F.; Masi, F.; Mallucci, G.; Bergamaschi, R.; Colombo, E.; Pascale, A. Disclosing the Novel Protective Mechanisms of Ocrelizumab in Multiple Sclerosis: The Role of PKC Beta and Its Down-Stream Targets. Int. J. Mol. Sci. 2024, 25, 8923. [Google Scholar] [CrossRef]
  225. Lis, M.; Niedziela, N.; Adamczyk-Zostawa, J.; Zalejska-Fiolka, J.; Szczygieł, J.; Sowa, A.; Świętek, A.; Adamczyk-Sowa, M. Comparative Effects of Vitamin D Supplementation on Oxidative Stress in Relapsing-Remitting Multiple Sclerosis. Curr. Issues Mol. Biol. 2024, 46, 14119–14131. [Google Scholar] [CrossRef]
  226. Bhaskaran, S.; Kumar, G.; Thadathil, N.; Piekarz, K.M.; Mohammed, S.; Lopez, S.D.; Qaisar, R.; Walton, D.; Brown, J.L.; Murphy, A.; et al. Neuronal deletion of MnSOD in mice leads to demyelination, inflammation and progressive paralysis that mimics phenotypes associated with progressive multiple sclerosis. Redox. Biol. 2023, 59, 102550. [Google Scholar] [CrossRef]
  227. Singh, I.; Samuvel, D.J.; Choi, S.; Saxena, N.; Singh, A.K.; Won, J. Combination therapy of lovastatin and AMP-activated protein kinase activator improves mitochondrial and peroxisomal functions and clinical disease in experimental autoimmune encephalomyelitis model. Immunology 2018, 154, 434–451. [Google Scholar] [CrossRef]
  228. Grujicic, J.; Allen, A.R. MnSOD Mimetics in Therapy: Exploring Their Role in Combating Oxidative Stress-Related Diseases. Antioxidants 2024, 13, 1444. [Google Scholar] [CrossRef]
Figure 1. MnSOD and the maintenance of mitochondrial redox homeostasis.
Figure 1. MnSOD and the maintenance of mitochondrial redox homeostasis.
Antioxidants 14 00848 g001
Figure 2. Coordination environment of Mn2+ and Mn3+ in human MnSOD. Comparison of the MnSOD active site in its reduced (Mn2+) and oxidized (Mn3+) states. Mn2+ is coordinated by His26, His74, His163, Asp159, and WAT1, with WAT2 hydrogen-bonded to Tyr34. In the Mn3+ state, WAT2 directly coordinates the metal center, replacing the Tyr34 interaction to complete the six-ligand geometry.
Figure 2. Coordination environment of Mn2+ and Mn3+ in human MnSOD. Comparison of the MnSOD active site in its reduced (Mn2+) and oxidized (Mn3+) states. Mn2+ is coordinated by His26, His74, His163, Asp159, and WAT1, with WAT2 hydrogen-bonded to Tyr34. In the Mn3+ state, WAT2 directly coordinates the metal center, replacing the Tyr34 interaction to complete the six-ligand geometry.
Antioxidants 14 00848 g002
Figure 3. Regulation of MnSOD transcription and activity.
Figure 3. Regulation of MnSOD transcription and activity.
Antioxidants 14 00848 g003
Table 1. Stepwise mechanism of MnSOD catalysis: associated residues, pathways, and kinetics.
Table 1. Stepwise mechanism of MnSOD catalysis: associated residues, pathways, and kinetics.
StepPathwayReaction DescriptionKey Residues and FeaturesRate ConstantReferences
1BothO2•− reduces Mn3+ to Mn2+; hydroxide ligand is protonated to H2O by Gln143Inner sphere: His26, His74, His163, Asp159, WAT1; Gln143 donates protonk1McAdam et al., 1977 [26]; Abreu et al., 2005 [27]
2FastSecond O2•− oxidizes Mn2+ to Mn3+; produces H2O2 via reaction with Tyr34 and water ligand Outer sphere: Tyr34, WAT1; enzyme is regeneratedk2McAdam et al., 1977 [26]; Perry et al., 2009 [28]; Bonetta Valentino et al., 2022 [20]
3SlowSecond O2•− binds Mn2+ to form a reversible peroxide adduct that inhibits activityProduct-inhibited complex formationk3McAdam et al., 1977 [26]; Bonetta Valentino et al., 2022 [20]
4SlowTwo protons regenerate Mn3+ and release H2O2, restoring activityProton donors: water ligand and outer sphere residuesk4McAdam et al., 1977 [26]; Bonetta Valentino et al., 2022 [20]
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

Grujicic, J.; Allen, A.R. Manganese Superoxide Dismutase: Structure, Function, and Implications in Human Disease. Antioxidants 2025, 14, 848. https://doi.org/10.3390/antiox14070848

AMA Style

Grujicic J, Allen AR. Manganese Superoxide Dismutase: Structure, Function, and Implications in Human Disease. Antioxidants. 2025; 14(7):848. https://doi.org/10.3390/antiox14070848

Chicago/Turabian Style

Grujicic, Jovan, and Antiño R. Allen. 2025. "Manganese Superoxide Dismutase: Structure, Function, and Implications in Human Disease" Antioxidants 14, no. 7: 848. https://doi.org/10.3390/antiox14070848

APA Style

Grujicic, J., & Allen, A. R. (2025). Manganese Superoxide Dismutase: Structure, Function, and Implications in Human Disease. Antioxidants, 14(7), 848. https://doi.org/10.3390/antiox14070848

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