Next Article in Journal
Evaluation of a Dietary Grape Extract on Oxidative Status, Intestinal Morphology, Plasma Acute-Phase Proteins and Inflammation Parameters of Weaning Piglets at Various Points of Time
Previous Article in Journal
Effects of Coenzyme Q10 Supplementation on Oxidative Stress Markers, Inflammatory Markers, Lymphocyte Subpopulations, and Clinical Status in Dogs with Myxomatous Mitral Valve Disease
Previous Article in Special Issue
NRF2 Alters Mitochondrial Gene Expression in Neonate Mice Exposed to Hyperoxia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Normal and Pathological NRF2 Signalling in the Central Nervous System

by
Tony Heurtaux
1,2,*,
David S. Bouvier
2,3,4,
Alexandre Benani
5,
Sergio Helgueta Romero
1,2,
Katrin B. M. Frauenknecht
2,3,
Michel Mittelbronn
1,2,3,4,6 and
Lasse Sinkkonen
1
1
Department of Life Sciences and Medicine (DLSM), University of Luxembourg, 4367 Belvaux, Luxembourg
2
Luxembourg Center of Neuropathology (LCNP), 3555 Dudelange, Luxembourg
3
National Center of Pathology (NCP), Laboratoire National de Santé (LNS), 3555 Dudelange, Luxembourg
4
Luxembourg Centre of Systems Biomedicine (LCSB), University of Luxembourg, 4367 Belvaux, Luxembourg
5
Centre des Sciences du Goût et de l’Alimentation, AgroSup Dijon, CNRS, INRAE, Université Bourgogne Franche-Comté, 21000 Dijon, France
6
Luxembourg Institute of Health (LIH), 1526 Luxembourg, Luxembourg
*
Author to whom correspondence should be addressed.
Antioxidants 2022, 11(8), 1426; https://doi.org/10.3390/antiox11081426
Submission received: 31 May 2022 / Revised: 18 July 2022 / Accepted: 19 July 2022 / Published: 22 July 2022
(This article belongs to the Special Issue NRF2 in Health and Diseases)

Abstract

:
The nuclear factor erythroid 2-related factor 2 (NRF2) was originally described as a master regulator of antioxidant cellular response, but in the time since, numerous important biological functions linked to cell survival, cellular detoxification, metabolism, autophagy, proteostasis, inflammation, immunity, and differentiation have been attributed to this pleiotropic transcription factor that regulates hundreds of genes. After 40 years of in-depth research and key discoveries, NRF2 is now at the center of a vast regulatory network, revealing NRF2 signalling as increasingly complex. It is widely recognized that reactive oxygen species (ROS) play a key role in human physiological and pathological processes such as ageing, obesity, diabetes, cancer, and neurodegenerative diseases. The high oxygen consumption associated with high levels of free iron and oxidizable unsaturated lipids make the brain particularly vulnerable to oxidative stress. A good stability of NRF2 activity is thus crucial to maintain the redox balance and therefore brain homeostasis. In this review, we have gathered recent data about the contribution of the NRF2 pathway in the healthy brain as well as during metabolic diseases, cancer, ageing, and ageing-related neurodegenerative diseases. We also discuss promising therapeutic strategies and the need for better understanding of cell-type-specific functions of NRF2 in these different fields.

Graphical Abstract

1. Beneficial and Harmful Roles of ROS

The nuclear factor erythroid 2-related factor 2 (NRF2) transcription factor pathway is able to neutralize reactive oxygen species (ROS) in order to maintain the cellular redox balance. Before going into details of the NRF2 regulatory pathway, we address the beneficial but also deleterious effects of ROS in this first section. It seemed important to us to summarize some background information, such as: how are ROS produced? What are their physiological roles? What are their deleterious effects? How is oxidative stress generated?

1.1. Radicals and Other Reactive Species at a Glance

Molecular oxygen (O2) is fundamental for survival of all aerobic organisms. According to its electronic structure, molecular oxygen has two unpaired electrons, allowing for capturing two electrons, albeit only one at a time. This electronic reduction of O2 then leads to the formation of metabolites, called reactive oxygen species (ROS), classified in radicals, including superoxide anion (O2●−), hydroxyl (HO), alkoxyl (RO), peroxyl (ROO), lipid (L), lipid alkoxyl (LO), and lipid peroxyl (LOO), and nonradical species such as hydrogen peroxide (H2O2), hypochlorous acid (HOCl), ozone (O3), and lipid peroxide (LOOH).
Reactive oxygen species are very short-lived but reactive molecules. They are mainly produced at membranes of mitochondria and the endoplasmic reticulum, in peroxisomes as well as in the cytosol [1,2]. The mitochondrial respiratory chain is considered the main source of physiological ROS production. During respiration, around 1 to 2% of O2 is not entirely reduced to water, therefore leading to the production of superoxide anion O2●− [2,3] (Figure 1A). Furthermore, an alteration of the inner mitochondrial membrane potential and electron transport rate can lead to an increase in the production of superoxide anions. ROS can also be produced in the cytosol by a large panel of active oxidoreductases such as NADPH oxidases (NOX), cytochrome P450 (CYP) oxidase, cyclooxygenases (COX), lipoxygenases (LOX), and monoamine oxidases (MAO) [4,5]. The lifespan of such reactive species is rather short, as they are quickly transformed by antioxidant proteins (e.g., superoxide dismutases (SOD), glutathione peroxidases (GPX), catalase (CAT)) into H2O2 and finally into water (Figure 1B); however, intermediate species can also be produced. Fenton and Haber–Weiss reactions, in which ferrous iron (Fe2+) reacts with H2O2, cause the formation of HO, constituting one of the most reactive species. HO and hydroperoxyl (HOO) radicals play an important role in the lipid peroxidation process by attacking lipids containing carbon–carbon double bonds such as polyunsaturated fatty acids [6,7] and in the subsequent production of reactive unsaturated aldehydes (e.g., 4-hydroxynonenal (4-HNE) and malondialdehyde (MDA)) [6].
Like ROS, reactive nitrogen species (RNS) are divided into radical and nonradical species, including nitric oxide (NO), nitrogen dioxide (NO2), peroxynitrite (ONOO), and nitrous acid (HNO2). Nitric oxide is produced by the activity of nitric oxide synthases (NOS), which convert the amino acid L-arginine into L-citrulline and NO (Figure 1B). A powerful oxidant, named peroxynitrite, is the result of the reaction between superoxide anion and nitric oxide [8]. Finally, the production of reactive species can be linked not only to metabolic activities but also to environmental factors, such as pollutants (pesticides, heavy metals), UV radiation, and lifestyle behaviours (tobacco smoke, excess alcohol, food habits, unbalanced exercise) [9,10,11,12].

1.2. Significance of ROS

Depending on their levels, ROS can be either beneficial or harmful to living systems. Over the past thirty years, many studies highlighted that ROS, at low concentrations, are important mediators of numerous signalling pathways. In particular, hydrogen peroxide, the most stable ROS with a relatively long biological lifespan (cellular half-life of 10−3 s; 1000 times more than other ROS), has the ability to diffuse through membranes, thus acting as an autocrine and paracrine signal [13,14]. Reactive species play useful biological roles. Indeed, ROS have been shown to be implicated in numerous physiological functions in embryonic and foetal development [15,16], neuronal development and function [17], and cellular proliferation and differentiation [18] due to notably supporting stem cell renewal and differentiation [19,20]. Roles in immune responses [21], activation of cell survival signalling pathways [22], normal growth and metabolism [23], tissue regeneration [24], blood pressure control [25,26,27], ageing prevention [28], and the execution of cell death programs such as apoptosis [22,29] can also be added to the long list of physiological functions attributed to ROS. Closely related to the synthesis of key signalling molecules, called eicosanoids (leukotrienes, lipoxins, prostaglandins, thromboxanes), lipid (hydro)peroxides are also considered as actors in normal physiological processes [6,30,31]. This eicosanoid synthesis is linked to COX and LOX activities. Importantly, these enzymes require [31,32] low levels of pre-existing lipid hydroperoxides (LOOH) to “prime” their catalytic cycles. Eicosanoids are lipid-signalling molecules deriving from arachidonic acid. These lipid mediators are key regulators of a wide variety of physiological responses as well as pathological processes. Many important cellular actions (e.g., proliferation, metabolism, and migration) are controlled by eicosanoids in the whole body [33,34,35]. An imbalance of this major lipid-signalling pathway contribute to disease progression [36]. It should also be noted that in the absence of ROS, alteration of the previously described functions will lead to decreased cell growth, metabolism, and proliferation as well as defective host defences [23].
Upon increased ROS levels, mammalian cells, equipped with a complex antioxidant defence system, will first try to maintain the redox balance [4,37,38]. For this purpose, enzymatic antioxidants (e.g., SOD, CAT, GPX), non-enzymatic scavengers (e.g., glutathione (GSH), thioredoxin, ascorbic acid (vitamin C), α-tocopherol (vitamin E), carotenoids), and also enzymes that repair cell damage (e.g., polymerases, nucleases, proteases) have a fundamental role in cellular protection. Redox balance, also called redox homeostasis, is an equilibrium between the production of ROS and these different detoxification systems (Figure 1C). However, an excessive and permanent increase in ROS production will induce an oxidative stress, which is classically defined as the alteration of the intracellular redox balance in favour of oxidative conditions. When ROS generation exceeds cellular capacity for detoxification, oxidative damage to nucleic acids (strand breaks, base oxidation), proteins (oxidation, nitration, carbonylation), lipids (peroxidation), membranes, and organelles (altered structures and properties) can occur, leading to cell dysfunction and cell death. Ageing and many human diseases such as cancer, diabetes, neurodegenerative disorders, cardiovascular diseases, and inflammation-related diseases have been linked to this progressive physiological dysfunction.

1.3. Redox Homeostasis in Ageing

According to the World Health Organization (WHO), the worldwide population over 60 years old will almost double, reaching 22% of the total population, by 2050. The proportion of people 80 years old or older is expected to triple between 2020 and 2050.
Ageing is a decline in physiological functions leading to the progressive loss of function in tissues and organs over time [39,40]. In 1956, the chemist Denham Harman published the “Free Radicals Theory of Ageing” [41]. He postulated that free radicals play a role in the ageing process. Although this hypothesis has been questioned many times, the fact that healthy ageing may be linked to oxidative stress resistance appears to still be relevant. The exact mechanism of oxidative-stress-induced ageing is still not clearly defined. However, it has been well established and accepted by the scientific community that a disrupted antioxidant/oxidant equilibrium coupled with an accumulation of oxidative damage to cellular constituents (DNA, proteins, lipids) and a chronic inflammation are pervasive features of ageing [42,43]. Key inflammatory players are indeed involved in the age-related process: activation of the pro-inflammatory nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) pathway, upregulation of cytokines (interleukin 1 beta (IL-1β), IL-6, tumour necrosis factor alpha (TNF-α)), C-C motif chemokine ligands (CCL2, CCL20) and their receptors, and other proinflammatory factors (matrix metalloproteinases, COX2, and NOS2) [42,44,45,46,47,48]. All these pro-inflammatory actors are indeed more expressed in aged than in young tissues [44,49]. The antioxidant/oxidant equilibrium is then disrupted during ageing, mainly due to a decrease in GSH synthesis as well as a downregulation of antioxidant enzymes [50,51]. A persistent oxidative stress associated with chronic inflammation may trigger ageing processes but also lead to age-related chronic diseases.

2. The NRF2-KEAP1 (Kelch-Like ECH-Associated Protein 1)-ARE (Antioxidant Response Element) Signalling Pathway

2.1. Mode of Regulation

As discussed in the previous section, the control of ROS levels is crucial to maintain cellular homeostasis. NRF2 is a transcription factor that emerged as a master regulator of the cellular antioxidant response.
Encoded by the NFE2L2 gene, NRF2 is a CNC-bZIP (Cap’n’Collar basic-region leucine zipper) transcription factor ubiquitously expressed in the body. Under normal conditions, NRF2 is kept in the cytoplasm by the KEAP1/CUL3-RBX1 E3 ubiquitin ligase complex, which is known to be the major repressor of NRF2 [52] (Figure 2). Beyond KEAP1, other regulators also exist. Indeed, BTB and CNC homology transcription factors, called BACH1 and BACH2, function as NRF2 repressors by competing with NRF2 binding to the ARE sequences [53,54,55,56]. BACH1 has been described to be widely expressed in mammalian tissues [54], whereas BACH2 is predominantly expressed in B and T lymphocytes, T cells, macrophages, and neural cells [57,58,59].
Sequestered in the cytosol, NRF2 is inactive and rapidly subjected to ubiquitination and proteasomal degradation (Figure 2). NRF2 is a protein with high turn-over. Complexed with KEAP1, the NRF2 half-life (t1/2) is approximately 20 min [60]. Oxidative conditions, following exposure to reactive chemicals, chemopreventive molecules (electrophiles agents), or an oxidative stress, will lead to the modification of two cysteine residues (C273 and C288) localized in the intervening region of KEAP1. As a consequence, the KEAP1-NRF2 complex will be disrupted, and NRF2 will be released (Figure 3). Several works have confirmed the role of such cysteine residues: mutations on cysteine 273 or cysteine 288 inactivate KEAP1 and promote the release of NRF2 [61,62,63]. Once free, stabilized NRF2 translocates into the nucleus, heterodimerizes with small musculoaponeurotic fibrosarcoma (sMAF) proteins, and binds to specific DNA sequences called antioxidant or electrophile response elements (ARE/EpRE, 5′-TGACNNNGC-3′) located in the 5′-flanking regions of gene promoters (Figure 3).
Post-translational modifications may also regulate NRF2 activities independently of KEAP1 expression [64]. Thus, the mitogen-activated protein (MAP) kinases, extracellular signal-regulated protein kinase (ERK) and c-jun N-terminal kinase (JNK), but also serine/threonine kinase (Akt), phosphatidylinositol 3-kinase (PI3K), 5′-AMP-activated protein kinase (AMPK), and protein kinase C (PKC) are able to phosphorylate the threonine, tyrosine and serine residues of NRF2, enhancing its activity [65,66,67,68,69]. In contrast, the glycogen synthase kinase-3 beta (GSK-3β) functions as an NRF2 inhibitor [70,71]. KEAP1 can also be subject to modifications by phosphorylation. Huo and collaborators described that the epidermal growth factor receptor (EGFR) tyrosine kinase was able to activate the NRF2 signalling pathway by phosphorylating KEAP1, thus resulting in nuclear NRF2 stabilization [72].

2.2. Biological Functions

NRF2 positively regulates hundreds of genes containing ARE sequences in their promoter regions [73]. The key well-known NRF2 function is linked to the maintenance of the redox balance through the synthesis or the use of γ-glutamyl-cysteinyl-glycine tripeptide (GSH). Thus, NRF2 is implicated in the de novo synthesis of GSH by the increase in glutamate-cysteine ligase (GCL) expression, the transformation of oxidized glutathione (GSSG) in GSH (glutathione reductase, GR), the expression of several enzymes using GSH for reducing peroxides (GPXs; peroxiredoxins), the breakdown of peroxides (CAT), and the import of cysteine (xCT transporter) important for GSH production (Figure 3).
NRF2 is also involved in the upregulation of the expression of drug/xenobiotic-metabolizing enzymes. Phase I and II drug-metabolizing enzymes play important detoxification roles by transforming the “parent drugs” into metabolites, which are less lipophilic and easier to eliminate through urine. NRF2 also modulates the transcription of Phase III drug-metabolizing enzymes implicated in the excretion of drugs/xenobiotics and metabolites from the cell. This will lead to a decrease in the therapeutic efficacy of drugs through their metabolism as well as through their elimination.
Numerous papers have also described the role of NRF2 in the reprogramming of cellular metabolism to support antioxidant responses [74,75]. Indeed, ARE sequences have been identified in genes implicated in glucose metabolism (glycolysis, pentose phosphate pathway, nucleotide biosynthesis pathway), lipid metabolism (lipid export, import, and synthesis), and heme and iron metabolism due to degrading free heme into biliverdin, CO, and Fe2+ (heme oxygenase 1, HMOX1), storing iron in its oxidized state (ferritin heavy chain 1, FTH1; ferritin light chain, FTL), and exporting unstable iron (ferroportin 1, FPN1) [53,76].
Autophagy processes and proteasome assembly are also governed by NRF2 transcription factor [77,78]. Autophagy plays an important role in removing misfolded or aggregated proteins as well as clearing damaged organelles (endoplasmic reticulum, mitochondria, and peroxisomes). Under normal conditions, the autophagy adaptor protein p62 binds to ubiquitylated protein aggregates in the cytoplasm before delivering them to the autophagosomes for degradation. When autophagy is disrupted, p62 accumulates in the cytoplasm and directly interacts with the NRF2-binding site on KEAP1. In this condition, the p62-KEAP1 complex is retained in the phagosomes, resulting in NRF2 stabilization and subsequently in transcriptional activation of NRF2 target genes [79,80]. Moreover, it has been described that ARE sequences are located on p62 promoter [53,81], indicating that NRF2 can also upregulate its transcriptional expression. In this way, p62 positively regulates NRF2 function independently of the cellular oxidative status. In contrast, a knockdown of p62 will decrease KEAP1 degradation and therefore NRF2 activation [82,83].
NRF2 levels vary significantly depending on physiological and pathological context. In recent years, many methods have been developed to monitor these changes in NRF2 pathway activity (Table 1), the aim being to assess NRF2 involvement in the pathogenesis of human diseases, to track disease progression but also to improve preclinical identification of targets, chemicals, and drugs.

3. Oxidative Stress, Inflammation, and NRF2 in the Brain

3.1. The Brain, an Ideal Target for Oxidative Attacks

The brain is particularly vulnerable to oxidative stress. To support intensive neuronal activity, the brain consumes almost 20% of the total O2 intake. This high O2 consumption associated with a high content of free metals (e.g., iron, copper, zinc, manganese), oxidizable polyunsaturated fatty acids, and auto-oxidation of neurotransmitters (dopamine, norepinephrine, and serotonin) may increase the risk of producing excessive ROS and, as a result, will make the brain vulnerable to ROS-mediated injury [102,103]. Compared to other organs (liver, kidney), the brain has fewer antioxidant defences, with lower CAT, GPX, and SOD activities. As an example, CAT concentration in the brain is 50 times lower than in the liver [104]. Disparities also exist within the brain itself. Unlike neurons, astrocytes and microglia appear to contain high GSH levels [105]. Neuronal GSH level is at least 50% lower than in other cells [103]. The major antioxidant scavenger, GSH, which is a γ-glutamyl-cysteinyl-glycine tripeptide, is important for cellular defence against reactive oxygen/nitrogen species. In the presence of reactive species, two molecules of GSH are oxidized to one molecule of GSSG, which is afterwards reduced to GSH by glutathione reductase. Astrocytes can convert glutamate released by neurons into glutamine and γ-glutamyl-cysteine, two GSH precursors. Since neuronal GSH levels are modest, astrocytes provide precursors (glutamine, cysteinyl-glycine dipeptide) to neurons to produce GSH themselves [106] to efficiently degrade H2O2. In addition to GSH release, astrocytes have also been described to produce other antioxidant molecules (e.g., ascorbate and vitamin E) and to activate ROS-detoxifying enzymes to protect neurons [105]. During oxidative insults, both astrocytes and microglia focus on maintaining the brain homeostasis in order to protect neurons. However, a huge production of reactive species as well as a GSH deficiency will contribute to a severe oxidative stress and will then play a role in the onset/aggravation of related pathological diseases [107,108].

3.2. NRF2, Driver for a Healthy Brain?

Concerning the brain, numerous studies have described NRF2 as a transcription factor mainly expressed in glial cells [109,110,111,112]. Considering the Brain RNAseq data from Barres’ Lab (www.brainrnaseq.org accessed on 20 June 2022), NFE2L2 expression level is indeed stronger in glial cells (microglia, astrocytes) compared to neuronal cells in both mouse and human. KEAP1, the major repressor of NRF2, appears to be expressed in all cells at a substantially identical level.
Over time, the relationship between NRF2 and brain health has been clearly established. The transcription factor has significant neuroprotective effects in brain diseases and injuries. As evidence, several studies have suggested that NRF2-deficient animals are more prone to inflammatory, cytotoxic, genotoxic, and neurotoxic effects due to endogenous and exogenous stressors [113,114,115,116,117]. Johnson and collaborators demonstrated that the absence of NRF2 in an experimental autoimmune encephalomyelitis model exacerbates the development of the disease, resulting in a more severe clinical course, a faster onset, and a higher percentage of diseased mice [118]. An NRF2 deficiency also worsened inflammatory parameters in a mouse model with combined tauopathy and amyloidopathy [119], increased Alzheimer’s disease (AD)-mediated cognitive decline [120], and significantly attenuated the self-renewal of glioma stem cells [121]. Furthermore, NRF2 deficiency mimics the ageing phenotype by exacerbating neuroinflammation, obesity-induced oxidative stress, blood-brain-barrier disruption, and cognitive decline in mice [122]. These observations coming from animal experiments have been confirmed in recent human studies [123,124,125,126,127,128].

3.3. Anti-Inflammatory Contribution of NRF2

Inflammation is a complex process whose final goal is to repair tissues by reducing and eliminating the damage caused. Age, genetics, lifestyle (e.g., smoking and diet), and environmental pollutants are factors that can promote acute or chronic inflammation [129]. While acute inflammation is time-limited and beneficial to the host, chronic inflammation is described as a slow, silent, and long-term process increasing health threat to the host. Chronic inflammation is a common feature of many pathological conditions such as asthma, allergy, diabetes, inflammatory bowel diseases (ulcerative colitis and Crohn’s disease), cancer, obesity, rheumatoid arthritis, and neurodegenerative diseases.
The NF-κB transcription factor regulates numerous genes involved in different processes of the immune and inflammatory responses in the whole body [130]. A dysregulated NF-κB pathway is thus considered as a hallmark of chronic inflammatory diseases. Classically, the NF-κB complex is a homo- or heterodimer composed of subunits such as p65 (also known as RelA), RelB, c-Rel, p100 (p52), and p105 (p50) [131]. In basal conditions, inactive NF-κB is associated with the inhibitory protein IκBα and sequestered in the cytoplasm. Under inflammatory/oxidative conditions, IκBα is phosphorylated by an IκBα kinase (IKK), leading to the dissociation of the NF-κB/IκBα complex and to the translocation of NF-κB into the nucleus, where it binds to specific response elements and upregulates the transcription process. The activated NF-κB cascade notably promotes the secretion of pro-inflammatory mediators (e.g., cytokines, chemokines), which will then recruit immune cells, resulting in the production of ROS and RNS.
A functional NRF2 system is important to regulate both neuroinflammation, i.e., activation of microglia and astrocytes, and oxidative stress in the brain. NRF2 and NF-κB transcription factors regulate cellular responses to inflammation and oxidative stress in order to maintain brain homeostasis [123,132,133]. Both pathways have been described to inhibit each other [134]. The anti-inflammatory role of NRF2 is now well established [135,136,137]. NRF2 inhibits the activation of NF-κB pathway by increasing antioxidant defences as well as HO-1 expression. The primary role of HO-1 is to reduce the pro-oxidant heme levels. Furthermore, the byproducts of HO-1 activity, carbon monoxide (CO, anti-apoptotic), ferrous iron (Fe2+), and bilirubin (anti-inflammatory), mitigate deleterious effects of oxidative stress by neutralizing ROS and detoxifying toxic chemicals [138,139,140,141]. The NRF2 pathway has also been described to inhibit NF-κB signalling through a KEAP1-induced IKK proteasomal degradation mechanism [133,142,143]. When NF-κB and the NRF2 pathways are simultaneously activated, NF-κB-p65 subunit may repress the transcriptional activity of NRF2. Recent works indeed have detailed that NF-κB can interact with KEAP1 in the cytoplasm. The formation of this complex promotes the entry of KEAP1 in the nucleus and facilitates the dissociation of NRF2 from the ARE sequences. The KEAP1-NRF2 complex thus formed will be exported from the nucleus and will then be degraded [144,145]. In response to oxidative stress, the NF-κB can also silence the NRF2-ARE pathway through the recruitment of the histone deacetylase3 (HDAC3), MAFK hypoacetylation [146,147,148], and a direct interaction between HDAC3 and MAFK, preventing any interaction with NRF2 and thus any NRF2 transcriptional facilitation [149]. Based on NRF2-deficient mouse experiments, a more pronounced NF-κB activity is observed, leading to a low inducible HO-1 expression, ROS accumulation, and increased inflammatory response [150,151,152].

4. NRF2 and Brain (Dys)Functions

4.1. NRF2 in Brain Control of Energy Metabolism

NRF2 has been described to play a role in obesity and metabolism [153] (Table 2). Following cellular stresses, the NRF2 pathway is activated to stimulate defence mechanisms, reduce body weight, and increase energy expenditure. Jardim and colleagues recently described that young mice, exposed to a high-caloric diet during their early-life period, had learning and memory impairments [154]. The authors revealed that this specific diet induced an oxidative stress in the hippocampus, a brain region known to have critical functions in cognition and memory but also to play certain roles in regulating food intake [155]. A high-caloric diet reduces NRF2 recruitment, whereas a dietary energy restriction triggers its activation and thereby induces many beneficial effects on health [156,157].
Other recent studies show that NRF2 is involved in brain control of food intake [158]. The hepatokine FGF21 is an endocrine signal, whose blood level increases during fasting, protein restriction, or high carbohydrate intake [159,160,161,162]. At the molecular level, FGF21 can stimulate the nuclear translocation of NRF2 and its recruitment to specific DNA regions that contains ARE sequences, triggering the transcription of the Oxt gene in oxytocin (OXT) neurons. These OXT neurons promote fat intake and limit sugar intake [158]. Thus, the FGF21-NRF2-oxytocin axis is viewed as an integrative system that contributes to the macronutrient-based diet selection, shifting from sugar to fat preference as an adaptive response to changes in the metabolic state. Parallel FGF21-sensitive glutamatergic neuronal pathways coexist to suppress carbohydrate intake independently of OXT neurons [163]. Through upregulation of NRF2 levels, an FGF21 administration may also show positive effects by protecting against diabetes-induced blood-brain-barrier disruption [164,165]. Inversely, in periphery, NRF2 regulates hepatocyte FGF21 secretion [166,167]. Interestingly, activation of NRF2 in hepatocytes in obese mice represents a strategy to stimulate FGF21 release in the blood, which decreases lipogenesis in the liver, promotes lipolysis in the white adipose tissue, and stimulates thermogenesis in the brown adipose tissue, thus preventing obesity and associated hepatosteatosis [167].
Table 2. NRF2 and brain control of energy metabolism.
Table 2. NRF2 and brain control of energy metabolism.
CharacteristicsNRF2 Pathway StatusKey PointsReferences
Fasting (in mice)downregulation in the brainreduced FGF21 sensitivity, reduced oxytocin signaling, increased sugar preference[158]
Sugar intake (in mice)activation in the brainincreased FGF21 sensitivity, activated oxytocin signaling, reduced sugar preference [158]
Diabetes (in db/db diabetic mice)downregulation in the braincorrelation with blood-brain-barrier permeability[164,165]
downregulation in the brain reversed by FGF21 treament blood-brain-barrier permeability reversed by FGF21 treament [165]
Diabetes, obesity (TrspRipKO mice)downregulation in the brainreduction in anorectic POMC neurons, loss of leptin sensitivity[168]
downregulation in the brain reversed by Keap1 invalidationreduction in anorectic POMC neurons, loss of leptin sensitivity reversed by Keap1 invalidation[168]
The role of NRF2 in control of brain energy homeostasis has been further demonstrated in diabetic animals, in which diabetes has been induced by genetic Trsp deficiency [168]. The Trsp gene encodes the tRNA for selenocysteine (tRNAsec), which is essential for the biosynthesis of selenoproteins, a group of proteins that contain selenocysteine residues [169], such as GPX and thioredoxin reductase. Trsp deletion in hypothalamic cells depletes selenoproteins in the hypothalamus, which provokes a local oxidative stress. This cellular stress alters and probably kills anorectic pro-opiomelanocortin (POMC) neurons, reducing leptin sensitivity in the hypothalamus, a crucial step in the pathogenesis of the metabolic disorder [168]. Induction of NRF2 in this model prevents oxidative damage in POMC neurons and reduces metabolic impairments.
These studies suggest that stimulation of NRF2 activity in the brain is a way to ameliorate whole-body energy homeostasis. Interestingly, NRF2 activity can be modified by daily consumption of specific food and phytochemicals. Most of these micronutrients/compounds are contained in plant-based food such as kaempferol (a polyphenol found in some fruits and vegetables), niga-ichigoside (a triterpenoid saponin from rosacea), salacia chinensis extracts (a climbing plant that contains triterpenes, phenolic compounds, and glycosides), hydroxytyrosol (the main phenolic compound in olives), curcumin, rebaudioside (a sweetener from Stevia rebaudiana), chicoric acid, and omega 3 and polyunsaturated fatty acids [170,171,172,173,174,175,176,177,178]. In addition, NFE2L2 gene expression can be upregulated by fibre-enriched diets [179]. Effects of these nutritional interventions on NRF2 were found in peripheral organs, and action on brain remains to be determined. It also remains to be established whether effects of these food-derived compounds on NRF2 are direct and/or associated with a better overall health.

4.2. NRF2 in Brain Ageing

Humans are among the longest-living mammals, with a maximum lifespan potential (MLSP) of 100 years. In animal models, the naked mole rat represents a unique model of healthy ageing, with a MLSP of 31 years [180]. NRF2 activity correlates with the MLSP of these naked mole rats. Indeed, it has been reported that long-living animal species have higher NRF2-signalling levels, highlighting the importance of NRF2 protection against ageing and ageing-related diseases [180].
In normal ageing, NRF2 activity, as well as expression of NRF2 target genes, decrease with age in many organs, including the brain [181,182] (Table 3). Consequently, the antioxidant/oxidant equilibrium is disrupted during brain ageing, thus favouring oxidative stress and associated damage. From a general point of view, the age-related progressive decline in NRF2 activity has been attributed to decreased expression of its positive regulators (PI3K, p62, CBP), increased expression of its negative regulators (KEAP1, BACH1), and also a general decrease in NRF2 protein expression itself [183,184,185]. As a result of NRF2 activity decline during ageing, GSH content as well as GSH- and antioxidant-related enzyme activity decrease in rat brain, while mitochondrial production of superoxide anions gradually increases over time [186,187]. These data obtained from animal experiments have been confirmed in a few human cohort studies, in which oxidative stress appears to be strongly linked to cognitive decline during ageing [188,189,190,191].
Multiple studies show that NRF2, whose activity is reduced during ageing, can be considered as a targetable pathway [181,182,192,193]. Thus, the activation of NRF2 in ageing could theoretically be beneficial. However, one recent study reports that the effect of knocking out NRF2 in old mice is rather contradictory to this conclusion [194]. Han and colleagues described that 18-month-old NRF2-KO mice had significantly better motor skills, less apoptosis in substantia nigra (SN), more tyrosine hydroxylase (TH)-positive signals in the striatum, a reduction in the 4-HNE staining in dopaminergic neurons of SN, and decreased iron accumulation in various brain regions compared to WT mice. Han and colleagues claimed that the regulation of brain iron metabolism in these old NRF2-KO mice was mediated by a decrease in FPN1 level, the only known exporter of non-heme iron in mammals, in brain endothelial cells, thus limiting the entry of iron into the brain and the subsequent formation of harmful reactive species, such as HO. This NRF2-KO mouse model, however, presents collateral damages with an exacerbation of the dysregulation of iron metabolism, an increase in ROS, and higher MDA levels in liver and spleen [195]. The results obtained by Han and colleagues contradict other studies. Indeed, the downregulation of ferroportin 1 in hippocampus of AD mouse models and also AD patients is associated with brain atrophy and cognitive impairment [196]. More recently, ferroportin loss in ageing brains has been described to be a key reason for iron mismanagement, and ferritin accumulation induced protein aggregation in Down syndrome, dementia, Alzheimer’s disease, and Parkinson’s disease (PD) [197]. The above studies claim that restoring Fpn expression, an NRF2 target gene coding for FPN1, in ageing and neurodegenerative models will ultimately result in a decrease in ferroptosis, protein aggregation, and memory loss.
Table 3. NRF2 and ageing.
Table 3. NRF2 and ageing.
CharacteristicsNRF2 Pathway StatusKey PointsReferences
Normal ageingdecrease with age in many organs, including the braindisruption of the antioxidant/oxidant equilibrium[181,182]
Normal ageingincreased activity in liver of long-lived rodentsincreased NRF2:ARE binding and antioxidant enzymatic activities[183]
Normal ageingdecrease with age in the braindecline in antioxidant enzyme activity after 12 months and incraeased superoxide anion in mitochondria[186,187]
Normal ageingpossibly decreased based on blood levels of lipoperoxidationindication of higher oxidative stress associated with cognitive decline[188]
Normal ageingpossibly decreased based on blood levels of antioxidant enzymeindication of lower antioxidative response associated with cognitive decline[189]
Normal ageingnot directly implicatedincreased oxidative stress associated with increased morbidity[190]
Normal ageingpossibly decreased based on blood levels of antioxidant enzymeindication of lower antioxidative response associated with cognitive decline[191]
Normal ageing upon NRF2 deletioninactive due to NRF2 knockout in mousereduced midbrain oxidative stress and reduced motor dysfunction in aged mice[194]
Normal ageing upon NRF2 deletioninactive due to NRF2 knockout in mouseincreased oxidative stress in liver and spleen[195]
Cell culture model of HGPSdecreased due to sequestering of NRF2 by progerinoxidative stress and aging defects in HGPS depend on NRF2 activity[198]
Premature ageing is a hallmark of Hutchinson–Gilford progeria syndrome (HGPS), a rare genetic disorder. De novo mutation in the LMNA gene induces the expression of progerin, a truncated lamin A protein that triggers various cellular defects. NRF2 has been recently identified to be one of the progerin’s targets [198]. Using human skin fibroblasts containing GFP-progerin, Kubben and colleagues revealed that progerin sequesters NRF2 and causes its mislocalisation to the nuclear periphery, altering its transcriptional activity and promoting oxidative stress. Interestingly, introducing constitutively active NRF2 to the HGPS fibroblast model restored most of the ageing-associated defects and showed the potential of the NRF2 pathway as an ageing modifier.
A deterioration of NRF2 activity is a critical point in the ageing process as well as age-related diseases. In order to restore the oxidative balance, many antioxidant supplementation therapies (e.g., antioxidant vitamins, coenzyme Q, resveratrol, curcumin) have been implemented. However, tests of the effectiveness of such treatments in humans have raised contrasting results [199,200,201]. Some NRF2-activating compounds have been described as potential senotherapeutic drugs. There are two kinds of senotherapeutics: senolytics, which induce cell death (also called senolysis) of senescent cells (e.g., curcumin, quercetin, berberine), and senomorphics, which suppress pro-inflammatory senescence-associated secretory phenotypes produced by senescent cells (e.g., rapamycin, metformin, epigallocatechin gallate) [202,203,204,205,206]. A promising avenue, increasingly raised in the literature, would then be to target the senescence process. Cellular senescence, one of the hallmarks of ageing, is characterized by the attainment of a quiescent state where cells lose their ability to divide and resist cell death. In the brain, the proportion of senescent glial and neuronal cells increases with age [207] and could be implicated in the onset or aggravation of cognitive impairment [208]. Many preclinical studies show positive effects of senolytics, which notably improve cognitive abilities in rodents and thus reduce the symptoms of age-related diseases [209,210]. Therefore, NRF2 can play a key role by regulating the multiple senescence-associated pathways [211]. The re-activation of NRF2 might thus be able to control and delay the ageing process. It remains to be evaluated how and when targeting NRF2 in ageing could be beneficial or detrimental. Key points regarding the role of NRF2 in ageing have been detailed below in Table 3.

4.3. NRF2 in Age-Related Neurodegenerative Disorders

In recent years, many findings describing the central role of the NRF2 transcription factor in redox homeostasis and anti-inflammatory functions in neurodegenerative disorders (NDDs) have emerged [212] (Table 4). As mentioned above (Section 3.2), brain NRF2 expression is found to be higher in glial cells (astrocytes and microglia) than in neurons, whereas the major NRF2 repressor, KEAP1, appears to be equally expressed in all brain cells. We can then postulate that glial cells are able to protect neurons against oxidative damage. As life expectancy is increasing, the prevalence of neurological diseases is likely to rise accordingly. Age-dependent cognitive loss as well as NRF2 activity decline are observed in NDDs such as AD and PD [213,214,215].
NRF2 appears to be an ideal target to modulate inflammation, autophagy, proteostasis, and oxidative stress in NDDs. Animal models helped to partially dissect the pathological mechanism in neurodegenerative diseases. As proof, the ablation of NRF2 in APP/PS1 mouse model leads to an increase in the beta-amyloid (Aβ) level (without increasing the number of Aβ plaques), pro-inflammatory molecules (production of IFNγ, IL-1β, IL-6, and TNF-α), oxidative DNA damage (increase in 8-hydroxydeoxyguanosine, 8-OHdG), glial cells reactivity (increase in Iba1 and GFAP staining), defective neuronal autophagy, and exacerbation of memory and cognitive defects in 11-to-12-month-old transgenic mice [120,216,217]. Knocking out NRF2 in a bigenic AD mouse model (AT-NRF2-KO) with combined amyloidopathy (expression of human mutant hAPPV717I) and tauopathy (expression of human mutant hTAUP301L protein) [119] also promoted inflammatory and oxidative stress. These AT-NRF2-KO mice have higher cognitive impairments and lower hippocampal long-term potentiation (lower electrophysical capacity of hippocampal neurons) already at six months of age before the appearance of Aβ and Tau pathologies.
The lentivirus-mediated expression of NRF2 in hippocampi of 9-month-old APP/PS1 mice led to an increase in insoluble Aβ, but it also led to attenuation of astrocyte reactivity, higher induction of HO-1 (neuroprotection), and reduction in spatial learning impairments [218]. The analysis of the astrocyte translatome, obtained by translating ribosome affinity purification sequencing (TRAP-seq), of the two AD mouse models, APP/PS1 and MAPTP301S mice, showed partial overlap of gene expression levels with age-altered pathways [219]. Jiwaji and colleagues crossed a GFAP-NRF2 mouse strain, mice overexpressing NRF2 in astrocytes, with APP/PS1 or MAPTP301S mice. NRF2 overexpression in APP/PS1 astrocytes reduced amyloid pathology as well as glial cells reactivity (decrease in Iba1 and GFAP expression) and reversed cognitive deficits observed in APP/PS1 mice. Cortical neurodegeneration and Tau pathology were decreased in MAPTP301S mice crossed with GFAP-NRF2 mice compared to the MAPTP301S mice. Some other studies report that NRF2 nuclear expression levels, in neurons and astrocytes, were decreased in AD and AD variant with Lewy bodies in CA1 of the hippocampus or cortical regions compared with age-matched control cases [120,220]. Thus, manipulation of astrocyte response through NRF2-associated pathways represents a promising therapeutic opportunity in AD.
An unbalance of NRF2 expression was demonstrated in autopsy samples of AD and PD patients. However, nuclear NRF2 staining was preserved, even amplified, in PD substantia nigra neurons [220,221] but insufficient to protect these neurons from degeneration. The expression of a PD-linked gene encoding α-synuclein (α-syn) in dopaminergic neurons of Drosophila [222], overexpression of human wild-type α-syn within murine SNpc [223], and toxin-based models by injection of 6-hydroxydopamine [224] and 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine [212,223,225,226] were used to mimic in vivo PD. In such models, the authors described a decreased locomotor activity, a progressive neuronal loss (TH-positive neurons), an intense astrogliosis and microgliosis, and a decrease in NRF2 activity.
Anandhan and collaborators described that NRF2 knockout in a PD mouse model (overexpression of human α-synuclein) resulted in an increase in the PD pathology associated with a behavioural dysfunction [126]. Indeed, at 3 months of age, hα-Syn+/NRF2-/- mice showed increased cognitive defects. At molecular and cellular levels, authors showed higher levels of phospho-α-synuclein, oxidation of lipids (increase in 4-HNE), increase in pro-inflammatory gene expressions (Nos2, Ptgs2), and increases in microglial activation and autophagy, especially in the midbrain and striatum. They also found a higher loss of tyrosine hydroxylase-positive neurons in nigrostriatal regions. An overexpression of human alpha-synuclein in the ventral midbrain of NRF2 knockout mice via the stereotaxic delivery of a viral vector leads to a more severe loss of dopaminergic neurons, inflammation, and microglia activation [227]. In this study, microglia were particularly altered toward pro-inflammatory/low phagocytotic profiles, showing again the important role of the induction of the NRF2 pathway in glial cell responses in NDDs.
All these results support new approaches in which targeting the NRF2 pathway could prevent and/or delay tissue injury. This has been notably confirmed by the use of celastrol. This natural bioactive ingredient derived from the Tripterygium wilfordii Hook F plant, described to activate the NRF2 pathway, decreases neuronal death, attenuates neuroinflammation, and relieves motor deficits [223] in PD mouse models. It was also described by Zhang and colleagues that celastrol could inhibit the NLRP3 inflammasome pathway, a critical component of the innate immune system, leading to caspase-1 activation and pro-inflammatory cytokines (IL-1β and IL-18) secretion in these PD mouse models. Many other studies have clearly shown that NRF2 overexpression led to beneficial effects in brain pathologies. Thus, astrocyte-specific overexpression of NRF2 delays motor deficits and synuclein aggregation in an A53T α-syn mutant mouse model [228], delays neurodegeneration in amyotrophic lateral sclerosis mouse models [229], and protects against optic tract damage and behavioural alteration in a mouse model of cerebral hypoperfusion [230]. Beneficial effects are also described in a mouse model with a fatal neurodegenerative disorder, called Alexander disease, caused by dominant mutations in the astrocyte intermediate filament glial fibrillary acidic protein leading to astrocyte overactivation [231]. All these preclinical data validated that NRF2 upregulation has neuroprotective effects.
Table 4. NRF2 and NDDs.
Table 4. NRF2 and NDDs.
CharacteristicsNRF2 Pathway StatusKey PointsReferences
AD mouse model APP/PS1genetic ablation of Nfe2l2increase in the amyloid level, oxidative and inflammatory markers, autopahgy, gliosis, and cognitive impairments[120,216,217]
AD mouse model APP/PS1lentivirus-mediated expression of NRF2 in hippocampiincrease in insoluble Abeta, decrease in astrogliosis, higher HO-1 level, reduction in spatial learning impairments[218]
AD mouse model bigenic APP/TAUgenetic ablation of Nfe2l2increase in oxidative and inflammatory markers, cognitive impairments, and lower LTP[119]
AD mouse model APP/PS1crossed with GFAP-NRF2, overexpression in astrocytesreduction in amyloid pathology, gliosis, and cognitive deficits[219]
AD mouse model MAPTP301Scrossed with GFAP-NRF2, overexpression in astrocytesreduction in Tau pathology, cortical neurodegeneration[219]
AD and AD variant with Lewy bodiesdecreased NRF2 levelshippocampus, cortex[120,220]
PDpreserved or amplified NRF2 levelssubstantia nigra neurons[220,221]
α-syn and toxins PD modelsdecreased NRF2 levelsgliosis, neuronal death, locomotor impairments[212,223,225,226]
α-syn PD mouse modelgenetic ablation of Nfe2l2higher levels of P-syn, inflammation, microgliosis, and autophagy, loss of TH neurons, increase in cognitive impairments[126]
A53T α-syn PD mouse modelcrossed with GFAP-NRF2, overexpression in astrocytesmotor deficits delayed[227]
Amyotrophic Lateral Sclerosis (ALS)crossed with GFAP-NRF2, overexpression in astrocytesneurodegeneration delayed[228]
Alexander’s diseaseNRF2 overexpression in astrocytesdecrease in GFAP expression and Rosenthal fibres, restoration of body weight[231]
Observations from post-mortem brains of PD patients reveal that oxidative stress may be implicated in the pathogenesis of PD [232,233]. This has been confirmed by an increase in oxidative damage markers (8-OHdG, MDA, 4-HNE) with decreased antioxidant defences in the brain and the blood of PD patients [234,235,236]. As a regulator of cellular antioxidant response, the NRF2 pathway appears as a potential target to prevent and/or delay tissue damage in PD. Indeed, results obtained from manipulation of the NRF2 pathway in animal models highlight the relevance of considering NRF2 as a potent modulator of neurodegeneration. Key points on the role of NRF2 in NDDs are resumed in Table 4.

4.4. NRF2 in Brain Cancer

Primary brain tumours, especially the heterogeneous group of diffuse gliomas, are still among the most devastating clinical conditions for which no cure exists in most cases. Understanding of how exactly those tumours either arise or infiltrate the brain is still incomplete to date. As the NRF2 pathway is involved in multiple key pathways in humans, it is worth having a closer look at its potential implication in brain tumorigenesis as well as its suitability as a treatment target. As oxidative stress is supposed to play a role in the development of high-grade gliomas, it was obvious that NRF2, as a central regulator of cell stress signalling, could be implicated (Table 5).
In the early days of brain tumour research related to the role of NRF2, it could be demonstrated in certainly oversimplified in vitro cell culture models that the downregulation of NRF2 led to increased levels of both apoptosis and autophagy [237,238]. The results obtained from more complex in vivo murine mouse models met those expectations by showing that reduced NRF2 levels, due to either direct downregulation of NRF2 or blocking its upstream activators ERK and PI3K, were associated with reduced glioma cell proliferation, enhanced apoptosis and ferroptosis, and impaired angiogenesis [239,240,241]. The anti-angiogenic effect of NRF2 could be linked to a decrease in hypoxia-inducible factor 1α (HIF-1α) stabilization, with a subsequent decrease in vascular endothelial growth factor (VEGF) expression [242]. A strong emphasis was also put on the positive effect that NRF2 exerted on the self-renewal capacity of glioma cells, as at that time, the maintenance of glioma stem cell properties was considered as a key factor in treatment resistance [121]. In the clinical context, NRF2 gained further attention, as it could be demonstrated that NRF2 expression was a negative prognostic factor for patient survival in glioma cohorts [243,244]. The negative association of NRF2 expression and glioblastoma patient survival could be later corroborated in a large cohort of The Cancer Genome Atlas (TCGA) and be linked mechanistically not only to its already described proliferation-enhancing properties but also to an invasion-stimulatory effect in the context of positive feedback between the autophagy-associated proteins SQSTM1/p62 and NRF2 [245]. Of note, gliomas with isocitrate dehydrogenase 1/2 (IDH1/2) mutations display significantly lower NRF2 levels as compared to their IDH1/2 wildtype diffuse glioma counterparts; however, in the subclass of IDH-mutant gliomas, high NRF2 levels were associated with a worse clinical prognosis [246,247]. IDH wildtype glioblastoma often display increased human telomerase reverse transcriptase (hTERT) activity in both hTERT wildtype and mutation contexts that can be stimulated by NRF2, thereby preventing ROS-induced glioma cell death [248].
A similarly detrimental clinical situation to the one of diffuse high-grade glioma is often encountered in patients with brain metastases. Most recently, it could be shown that even in early stages of non-small-cell lung cancer (NSCLC), high NRF2 expression levels in cancer cells were associated with an increased risk of developing brain metastases [249]. This NRF2 overexpression in more malignant clinical situations seems to be strongly related to different mutations in the NRF2-KEAP1-ARE cell survival pathway [250]. In most patients with NSCLC brain metastasis, those mutations were either specific or increased in brain metastases as compared to the primary tumours. The mutations either affect the capacity of KEAP1 to induce the degradation of NRF2 or alter the Neh2 domain of NRF2 that is responsible for binding KEAP1 constituting a prerequisite for subsequent degradation. Several mutations specifically led to NRF2 stabilization and nuclear translocation, with subsequent increased activation of its target genes. In the same study, the authors even went beyond studying NSCLC and detected different NRF2 mutations also in circulating tumour cells from patients with melanoma, breast, and colon carcinoma metastases [250]. In addition to the negative primary effects that could be attributed to NRF2 signalling, similarly detrimental functions were reported in the treatment context of brain cancer [251,252].
Table 5. NRF2 and cancer.
Table 5. NRF2 and cancer.
CharacteristicsNRF2 Pathway StatusKey PointsReferences
Glioma cell linedownregulationincreased apoptosis[237]
downregulationincreased autophagy[238]
overexpressionreversed ERK and PI3K-inhibition-induced inhibition of cell viability[239]
overexpressionincreased proliferation, resistance to ferroptosis, increased oncogenic potential[240]
downregulationreduced proliferation[241]
downregulationreduced proliferation, reduced mitochondrial oxygen consumption[242]
downregulationreduced proliferation[121]
upregulationincreased proliferation, tumour cell infiltration, and mesenchymal transition[245]
downregulationdecreased resistance towards chemotherapy[247]
overexpressionprevention of ROS-induced cell death[248]
downregulationenhanced sensitivity towards chemotherapy and irradiation[251]
overexpressionincreased cell survival after chemotherapy[252]
overexpressionreduced ROS levels, increased increased cell survival[253]
Human glioma tissueupregulation (as compared to normal tissue)potentially implicated in glioma progression[239]
upregulationworse patient survival[240]
upregulation (as compared to normal tissue)potentially implicated in glioma progression[241]
upregulationworse patient survival[241]
upregulationincreased angiogenesis[242]
upregulation with grade of malignancypotentially implicated in glioma progression[243]
cytoplasmic expressionassociation with worse prognosis[243]
nuclear expressionassociation with better prognosis[243]
overexpressionassociation with worse prognosis, more tissue necrosis, correlation with high HIF1alpha levels[244]
overexpressiondecreased progression-free survival[245]
overexpressionworse patient prognosis in anaplastic glioma[246]
overexpressionupon irradiation and chemotherapy[251]
overexpressionassociation with worse prognosis[252]
Xenograft glioblastomas mouse modeldownregulationreduced tumour growth, reduced proliferation, increased apoptosis, reduced angiogenesis[241]
downregulationreduced VEGF expression, reduced angiogenesis[242]
Human lung carcinoma tissueoverexpressionhigher risk of brain metastasis[249]
increased mutational rate of NRF2 pathwayhigher risk of brain metastasis[250]
Both chemotherapy and irradiation treatment enhanced NRF2 levels in human glioma cell lines and patient samples, and the degree of NRF2 expression was associated with earlier recurrences. A more recent study was able to show that NRF2-related protumorigenic properties in gliomas are not exclusively linked to ROS regulation or metabolic adaptation but are also linked to an enhanced expression of the transcriptional coactivator TAZ belonging to the HIPPO pathway, most probably by activating functional enhancers in the regulatory regions of TAZ [253]. As an upstream regulator of NRF2-mediated ROS tolerance in malignant glioma cells, the Wiskott–Aldrich syndrome protein (WASP)-interacting protein (WIP) inhibits KEAP1, which acts as an NRF2 repressor, finally leading to reduced ROS levels and unwanted glioma cell survival [253]. In summary, there is a large body of evidence that increased NRF2 levels are associated with (a) various tumour-promoting functions in malignant gliomas, (b) negative patient survival in diffuse gliomas and glioblastomas, (c) tumour cell senescence [254], and (d) resistance towards radio-chemotherapy.
Targeting NRF2 has mainly been restricted to preclinical cell culture and animal studies, while promising early clinical concepts are lacking. This might be related to the fact that, over the last several decades, pharmacological research regarding NRF2 modulating drugs has focused on NRF2-activating rather than -inhibiting drugs, with the primary goal to enhance cellular tolerance towards ROS, which might be beneficial in a plethora of degenerative disorders but is probably less promising as a novel concept in tumour treatment [255]. The major challenge in potential novel treatment approaches blocking NRF2 in cancer and especially in malignant gliomas will be to overcome expected severe negative side effects in residual cells, most probably neurons and muscle cells, which both may rapidly suffer from a decreased capacity to deal with high ROS levels, thereby being at risk to develop secondary neurodegenerative or myopathic symptoms.

5. Epigenetic Regulation of NRF2 Signalling

5.1. NRF2-Mediated Transactivation and Chromatin Level Regulation

The basic unit of chromatin is the nucleosome, which consists of eight histone proteins that are surrounded by ~147 bp fragments of DNA. These nucleosomes progressively condense to form chromatin [256]. Chromatin can be found in two main configurations: heterochromatin, which is highly compacted and has low accessibility for transcription factor binding; and euchromatin, which is more open and therefore where most gene activation and transcription occurs [257]. Promoters and enhancers are the regions of DNA that control gene expression. To initiate the transcription of a specific gene, it is necessary for transcription factors (TFs) to bind to the promoter of that gene, recruiting the rest of the machinery required for transcriptional activation (other TFs, cofactors, and RNA polymerase II holoenzyme (RNA Pol II)). Furthermore, enhancer elements are regulatory regions that interact with promoters and transcriptional machinery to stabilize the RNA Pol II complex and increase gene expression [258]. While the function of promoters depends on their position and orientation relative to the target gene, enhancers can interact with promoters located far away in chromatin due to three-dimensional loop formation [259,260].
Histone acetylation increases the accessibility of chromatin for transcription factor binding, which often links to gene transcription activation. On the other hand, deacetylation usually leads to gene repression. Histone acetyltransferases (HATs) and HDACs are the proteins that acetylate or deacetylate histones, respectively. Reduced NRF2 expression is associated with decreased histone acetylation at NFE2L2 promoter, and HDAC inhibition can increase NRF2 expression with beneficial effect in glial cells [261,262]. In contrast, histone methylation is associated with both gene repression and activation, depending on the residue that is methylated and the degree of methylation. Histone methyltransferases and histone demethylases are the proteins that methylate or demethylate the histones, respectively. Histone H3 lysine 27 trimethylation (H3K27me3) is a histone modification associated with Polycomb-mediated gene repression and found to be deposited at the NFE2L2 promoter by EZH2 enzyme of the PRC2 complex in lung cancer cells [263]. Thus, Polycomb-mediated repression of NFE2L2 might also occur in the different cell types of the brain.
Once NRF2 is released from KEAP1-mediated retention in the cytoplasm, it translocates to the nucleus and, after binding to ARE sequences on DNA, regulates the transcription of its target genes in collaboration with sMAF proteins (Figure 3). This process is called NRF2-mediated gene transactivation. ARE sequences are found in the promoters of NRF2 target genes, but there is also evidence of ARE sequences located at enhancer elements [264,265]. Typically, after NRF2 binds to the promoters of its target genes, recruitment of other TFs or cofactors is required. Some examples are FOS, MAFB, LHX3, MEF2A [265], ATF4 [266], or BRG1 [267]. It is also known that ARE-like sequences can be recognized by other Cap’n’Collar TFs, including BACH1 and BACH2 [56] and members of the AP1 complex [268]. A defining feature of TF-mediated transactivation is the recruitment of the Mediator complex that creates the physical interaction between the TF and RNA Pol II. In case of NRF2, the interaction takes place with a Mediator subunit called MED16 [269]. Moreover, there are studies showing that NRF2 can also associate with other transcriptional regulators such as chromodomain helicase DNA-binding protein 6 [270], receptor-associated co-activator 3 [271], and NAD+-dependent histone deacetylase sirtuin 6 [272]. In addition to protein-coding genes, NRF2 can also activate the expression of long non-coding RNAs after binding to their promoters [273].
The gene activation by NRF2 is accompanied by the recruitment of histone acetylases that increase local histone acetylation. This can be counteracted by activity of other factors such as NF-κB by depriving NRF2 of its co-activator and facilitating histone deacetylation via recruitment of histone deacetylases such as HDAC3 [149]. There is some evidence that NRF2 can repress the expression of a variety of target genes. This repression by NRF2 can be direct, through the formation of inhibitory complexes with other transcription factors [274], or indirect, for example, through the activation of the expression of microRNAs (miRNAs) [275]. Some examples of miRNAs that can be activated by NRF2 are miR-365-1/miR-193b cluster, miR-29-b1, and miR-125-b1 [264,276]. In line with this, studies such as the one by Kwak and collaborators indicate that NRF2 is primarily an activator and blocks gene expression indirectly [277].
The overall transcriptomic response to NRF2 activation depends on the presence of accessible ARE-containing enhancers available for NRF2 binding. The epigenetic mechanisms controlling enhancer accessibility at the chromatin level include DNA methylation, histone modifications, non-coding RNAs, and chromatin remodelling [278]. A large part of the accessible chromatin is determined by the cell type [279,280], but it is also determined by environmental exposures and genotype of the cell in question [281,282]. Through genome-wide chromatin immunoprecipitation analysis, thousands of NRF2 binding sites have been identified in vivo [283], with a large proportion of AREs showing cell-type-selective binding. As an interesting example, NRF2 binding sites at target genes associated with poor survival across multiple cancers have a particularly permissive chromatin structure across cell types, allowing broad activation of these genes [284]. Meanhile, other NRF2 targets, not associated with cancer, show more cell-type-selective chromatin accessibility at their ARE sites. Despite being a biomedically relevant target tissue, the occupied NRF2 binding sites in the different cell types of the central nervous system have not yet been extensively studied. However, as described below, this could yield important insights into the cell-type-specific activity of NRF2.
As mentioned above, the availability of AREs for NRF2-mediated activation is also controlled by competitive binding by BACH1 and BACH2 to the same sequence motif [53,54,55,56] (Figure 3). BACH1 and BACH2 are predominantly transcriptional repressors, thereby counteracting the NRF2-mediated activation [285,286] and possibly also inducing chromatin compaction at the AREs. Thus, the final transcriptomic response to NRF2 activation is determined by not only the array of available ARE-containing enhancers but also the balance between NRF2 and BACH1/2 expression levels in the target cell.

5.2. Different NRF2 Activity in Brain Cell Types and Approaches for Cell-Type-Specific Targeting

As for any gene, there are multiple determinants for NFE2L2 expression levels. Firstly, its promoter contains ARE sequences, allowing for auto-regulation through NRF2 binding [287] (Figure 3). In addition, a xenobiotic response element for aryl hydrocarbon receptor, NF-κB, and AP2 binding sites can be found at the promoter [287,288,289], among others. Moreover, KRAS signalling is known to induce NRF2 expression in the context of cancer [290].
Across brain cell types, NFE2L2 is the least expressed in different neurons, while high expression can be detected in non-neuronal cell types [111,291]. More specifically, the highest expression levels were detected in microglia and astrocytes, consistently with the above-discussed reliance of neurons on glial cells for their antioxidant response and neuroprotection [292,293,294]. The neuronal repression of NRF2 was found to be accompanied by histone H3 hypoacetylation at the NFE2L2 promoter and could be reversed by HDAC inhibitors in neurons [295]. The repression was induced during neuronal differentiation and maturation, leading to impaired antioxidant response, but could not be reversed by HDAC inhibition in very young neurons, suggesting an important developmental role. Indeed, the NRF2 repression was proposed to be necessary for normal neurodevelopment due to its interference with WNT signalling activation. Interestingly, in parallel with low expression of NRF2, midbrain dopaminergic neurons are enriched for high BACH2 levels [111], allowing further repression of NRF2 target sites in neurons and ensuring correct neuronal commitment. Given that NFE2L2 promoter hosts an ARE, it would be interesting to investigate whether deacetylation of the promoter is induced by increasing BACH2 binding during neurogenesis. Finally, in addition to transcriptional control, NFE2L2 mRNA was suggested to be targeted by multiple miRNAs in neuronal SH-SY5Y cell line, possibly further contributing to the low NRF2 expression in neurons [296].
While neurons are lacking in their own NRF2 signalling, glial cells and brain tumours derived from them, such as glioblastomas, show high NRF2 activity that is associated with a poor prognosis (Section 4.4). Therefore, depending on the pathology and the cell type affected, either further activation of NRF2 signalling, as in neurodegenerative diseases and ageing, or repression of NRF2 signalling, as in cancer, would be needed. For example, activation of NRF2 using compounds such as dimethyl fumarate can have beneficial effects in human and mouse models on both AD and PD [119,297,298,299,300,301,302]. Similarly, in some cancer models, an inhibition of NRF2 by small compounds has shown promise in in vitro assays [303]. However, the existing direct activators or inhibitors of NRF2 signalling often have high turnover and are difficult to deliver to the target tissue, creating further challenges for systematic treatments. Furthermore, as discussed above, systematic ablation of NRF2 activity can lead to side effects in other tissues such as spleen and liver [195].
The cancer-related hyperactivity of NRF2 signalling drives an overexpression of target genes located in permissive chromatin [284] but, in some cancers, such as non-small-cell lung cancer, also enhances reprogramming revealing novel cancer-specific ARE targets for NRF2 binding occurs [304]. Similarly, cell-type-specific profiles of ARE-containing enhancers are likely to exist in the different cell types of the brain and brain tumours [305,306,307]. Genome-wide mapping of ARE-carrying enhancers targeted by NRF2 and/or BACH1/2 in the different cell types of the brain and various brain tumours and understanding the contribution of those target genes to different pathologies might allow for more precise approaches for targeting NRF2 signalling. Indeed, analysis of BACH1 regulation in ventral midbrain found BACH1 target genes to be involved in several neuroprotective pathways but repressed during neurodegeneration [308]. Interestingly, inhibition of BACH1 enabled an NRF2-mediated increase in the target genes’ expression in neuronal cells and protected against loss of dopaminergic neurons in a mouse model of PD. Whether a similar approach could be taken with BACH2, which shows a more neuron-specific expression profile [111], would be interesting to test. In addition, other signalling pathways such as retinoic acid and oestrogen signalling are known to interfere with NRF2 signalling [309] and provide an alternative avenue for controlling the pathway. However, these approaches also require a more detailed understanding of the cell-type-specific target genes. With the arrival of low-input and single-cell technologies, rare cell types of the brain can now also be studied [310].
Finally, more detailed knowledge of the genome-wide action of TFs such as NRF2 and BACH1/2 is beneficial not only for better targeting of NRF2 signalling but also for diagnostics and better understanding of disease risk. Indeed, disease-associated regulatory genetic variants are known to be enriched in cell-type-selective enhancer regions, where they can disrupt or create TF bindings sites [311,312,313]. Through cell-type-specific alterations in gene expression, such variants can then influence disease risk and, when well understood, be used for diagnostic purposes. This is also true for enhancers bound by NRF2: Wang and collaborators identified several disease-associated single nucleotide polymorphisms capable of disrupting AREs and leading to allele-specific NRF2 binding [283]. Interestingly, a variant with strong association with parkinsonian disorders was found to decrease NRF2 binding at an ARE in the intron of the Microtubule Associated Protein Tau (MAPT) gene and to drive lower MAPT transcription in neuronal progenitors. This provides the first insight into the mechanistic explanation for this variant’s association with PD and related disorders.

6. Conclusions

Oxidative stress and inflammation are the two most common features of brain diseases. Here, we provided evidence that NRF2, by its anti-oxidant and anti-inflammatory properties, could be a promising therapeutic target for the maintenance of brain functions as well as for brain dysfunctions improvement.
Targeting the NRF2 pathway in brain diseases appears to be context- but also time-dependent (Figure 4). Generally, NRF2 recruitment plays a protective role under physiological conditions, but it also promotes the development of many cancers once cancer is established [314] (Figure 4A). In early stages of cancer, NRF2 is able to thwart the deleterious effects of ROS, such as oxidative damages caused to cellular constituents (DNA, proteins, lipids). Pölönen and colleagues indeed described different NRF2 activation levels depending on the glioma WHO grades I-IV [245]. They reported that NRF2 activity was gradually higher in grades II to IV, whereas no constitutive NRF2 activity was detected in grade I. NRF2 expression may therefore be a useful biomarker to predict WHO grade and cellular behaviour of brain tumours [315]. In more advanced cancer stages, where NRF2 levels are higher, the therapeutic procedure would then be to decrease NRF2 activation in order to limit angiogenesis, cancer cell growth, and both chemo- and radio-resistance of cancer cells [316].
As life expectancy is increasing, the prevalence of neurological diseases is likely to rise accordingly. Age-dependent cognitive loss is observed in neurodegenerative disorders such as Alzheimer’s and Parkinson’s diseases [213]. In parallel, NRF2 activity has also been reported to decrease with age [182,184,215] (Figure 4B,C). As we previously mentioned, NFE2L2 expression level is higher in glial cells (astrocytes, microglia) compared to neuronal cells in both mouse and human. Thus, discrepancies in targeting the NRF2 pathway could be brain-region- but also cell-type-dependent.
The future challenges will be to establish novel therapies to (a) specifically inhibit the NRF2 pathway in cancer cells, (b) increase NRF2 activation in specific cell types and/or brain regions, and (c) modulate NRF2 pathway in senescent cells. We believe that modulation of NRF2 signalling pathway by using specific food products and phytochemicals, dietary supplements, drugs, and epigenetic modifiers, alone or in combination, will help to limit inflammatory diseases, cancer progression, ageing process, and subsequently ageing-related diseases.

Author Contributions

All authors listed contributed to this work. T.H., D.S.B., A.B., S.H.R., K.B.M.F., M.M., and L.S. performed literature searches and wrote the manuscript. T.H. made the figures and built the structure of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

S.H.R. is supported by Fonds National de Recherche de Luxembourg (FNR)-PRIDE (PRIDE17/12244779/PARK-Q); L.S. has received funding from FNR, Fondation du Pélican de Mie et Pierre Hippert-Faber, Luxembourg Rotary Foundation, Luxembourg Personalized Medicine Consortium, and University of Luxembourg Internal Research Project grants; M.M. would like to thank the FNR for the support (PEARL P16/BM/11192868 grant).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no competing financial or non-financial interests.

References

  1. Di Meo, S.; Napolitano, G.; Venditti, P. Physiological and Pathological Role of ROS: Benefits and Limitations of Antioxidant Treatment. Int. J. Mol. Sci. 2019, 20, 4810. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Poyton, R.O.; Ball, K.A.; Castello, P.R. Mitochondrial generation of free radicals and hypoxic signaling. Trends Endocrinol. Metab. 2009, 20, 332–340. [Google Scholar] [CrossRef] [PubMed]
  3. Brand, M.D. The sites and topology of mitochondrial superoxide production. Exp. Gerontol. 2010, 45, 466–472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Snezhkina, A.V.; Kudryavtseva, A.V.; Kardymon, O.L.; Savvateeva, M.V.; Melnikova, N.V.; Krasnov, G.S.; Dmitriev, A.A. ROS Generation and Antioxidant Defense Systems in Normal and Malignant Cells. Oxid. Med. Cell. Longev. 2019, 2019, 6175804. [Google Scholar] [CrossRef]
  5. Panday, A.; Sahoo, M.K.; Osorio, D.; Batra, S. NADPH oxidases: An overview from structure to innate immunity-associated pathologies. Cell. Mol. Immunol. 2015, 12, 5–23. [Google Scholar] [CrossRef] [Green Version]
  6. Ayala, A.; Muñoz, M.F.; Argüelles, S. Lipid peroxidation: Production, metabolism, and signaling mechanisms of malondialdehyde and 4-hydroxy-2-nonenal. Oxid. Med. Cell. Longev. 2014, 2014, 360438. [Google Scholar] [CrossRef]
  7. Yin, H.; Xu, L.; Porter, N.A. Free radical lipid peroxidation: Mechanisms and analysis. Chem. Rev. 2011, 111, 5944–5972. [Google Scholar] [CrossRef]
  8. Pacher, P.; Beckman, J.S.; Liaudet, L. Nitric Oxide and Peroxynitrite in Health and Disease. Physiol. Rev. 2007, 87, 315–424. [Google Scholar] [CrossRef] [Green Version]
  9. Carraro, E.; Schilirò, T.; Biorci, F.; Romanazzi, V.; Degan, R.; Buonocore, D.; Verri, M.; Dossena, M.; Bonetta, S.; Gilli, G. Physical Activity, Lifestyle Factors and Oxidative Stress in Middle Age Healthy Subjects. Int. J. Environ. Res. Public. Health 2018, 15, 1152. [Google Scholar] [CrossRef] [Green Version]
  10. Di Meo, S.; Reed, T.T.; Venditti, P.; Victor, V.M. Role of ROS and RNS Sources in Physiological and Pathological Conditions. Oxid. Med. Cell. Longev. 2016, 2016, 1245049. [Google Scholar] [CrossRef]
  11. Al-Gubory, K.H. Environmental pollutants and lifestyle factors induce oxidative stress and poor prenatal development. Reprod. Biomed. Online 2014, 29, 17–31. [Google Scholar] [CrossRef] [Green Version]
  12. Aseervatham, G.S.B.; Sivasudha, T.; Jeyadevi, R.; Arul Ananth, D. Environmental factors and unhealthy lifestyle influence oxidative stress in humans—An overview. Environ. Sci. Pollut. Res. 2013, 20, 4356–4369. [Google Scholar] [CrossRef]
  13. Knaus, U.G. Oxidants in Physiological Processes. Handb. Exp. Pharmacol. 2021, 264, 27–47. [Google Scholar] [CrossRef]
  14. Bienert, G.P.; Schjoerring, J.K.; Jahn, T.P. Membrane transport of hydrogen peroxide. Biochim. Biophys. Acta Biomembr. 2006, 1758, 994–1003. [Google Scholar] [CrossRef] [Green Version]
  15. Jamil, M.; Debbarh, H.; Aboulmaouahib, S.; Aniq Filali, O.; Mounaji, K.; Zarqaoui, M.; Saadani, B.; Louanjli, N.; Cadi, R. Reactive oxygen species in reproduction: Harmful, essential or both? Zygote 2020, 28, 255–269. [Google Scholar] [CrossRef]
  16. Dennery, P.A. Oxidative stress in development: Nature or nurture? Free Radic. Biol. Med. 2010, 49, 1147–1151. [Google Scholar] [CrossRef]
  17. Oswald, M.C.W.; Garnham, N.; Sweeney, S.T.; Landgraf, M. Regulation of neuronal development and function by ROS. FEBS Lett. 2018, 592, 679–691. [Google Scholar] [CrossRef]
  18. Chaudhari, P.; Ye, Z.; Jang, Y.-Y. Roles of reactive oxygen species in the fate of stem cells. Antioxid. Redox Signal. 2014, 20, 1881–1890. [Google Scholar] [CrossRef] [Green Version]
  19. Morimoto, H.; Iwata, K.; Ogonuki, N.; Inoue, K.; Atsuo, O.; Kanatsu-Shinohara, M.; Morimoto, T.; Yabe-Nishimura, C.; Shinohara, T. ROS are required for mouse spermatogonial stem cell self-renewal. Cell Stem Cell 2013, 12, 774–786. [Google Scholar] [CrossRef] [Green Version]
  20. Sart, S.; Song, L.; Li, Y. Controlling Redox Status for Stem Cell Survival, Expansion, and Differentiation. Oxid. Med. Cell. Longev. 2015, 2015, 105135. [Google Scholar] [CrossRef] [Green Version]
  21. Bassoy, E.Y.; Walch, M.; Martinvalet, D. Reactive Oxygen Species: Do They Play a Role in Adaptive Immunity? Front. Immunol. 2021, 12, 755856. [Google Scholar] [CrossRef]
  22. Redza-Dutordoir, M.; Averill-Bates, D.A. Activation of apoptosis signalling pathways by reactive oxygen species. Biochim. Biophys. Acta Mol. Cell Res. 2016, 1863, 2977–2992. [Google Scholar] [CrossRef]
  23. Santos, A.L.; Sinha, S.; Lindner, A.B. The Good, the Bad, and the Ugly of ROS: New Insights on Aging and Aging-Related Diseases from Eukaryotic and Prokaryotic Model Organisms. Oxid. Med. Cell. Longev. 2018, 2018, 1941285. [Google Scholar] [CrossRef]
  24. Dhoke, N.R.; Geesala, R.; Das, A. Low Oxidative Stress-Mediated Proliferation Via JNK-FOXO3a-Catalase Signaling in Transplanted Adult Stem Cells Promotes Wound Tissue Regeneration. Antioxid. Redox Signal. 2018, 28, 1047–1065. [Google Scholar] [CrossRef] [Green Version]
  25. Lassègue, B.; Griendling, K.K. Reactive oxygen species in hypertension: An update. Am. J. Hypertens. 2004, 17, 852–860. [Google Scholar] [CrossRef] [Green Version]
  26. Kröller-Schön, S.; Daiber, A.; Steven, S.; Oelze, M.; Frenis, K.; Kalinovic, S.; Heimann, A.; Schmidt, F.P.; Pinto, A.; Kvandova, M.; et al. Crucial role for Nox2 and sleep deprivation in aircraft noise-induced vascular and cerebral oxidative stress, inflammation, and gene regulation. Eur. Heart J. 2018, 39, 3528–3539. [Google Scholar] [CrossRef]
  27. Steven, S.; Frenis, K.; Kalinovic, S.; Kvandova, M.; Oelze, M.; Helmstädter, J.; Hahad, O.; Filippou, K.; Kus, K.; Trevisan, C.; et al. Exacerbation of adverse cardiovascular effects of aircraft noise in an animal model of arterial hypertension. Redox Biol. 2020, 34, 101515. [Google Scholar] [CrossRef]
  28. Kong, H.; Chandel, N.S. Regulation of redox balance in cancer and T cells. J. Biol. Chem. 2018, 293, 7499–7507. [Google Scholar] [CrossRef]
  29. Circu, M.L.; Aw, T.Y. Reactive oxygen species, cellular redox systems, and apoptosis. Free Radic. Biol. Med. 2010, 48, 749–762. [Google Scholar] [CrossRef] [Green Version]
  30. Zhang, R.; Brennan, M.-L.; Shen, Z.; MacPherson, J.C.; Schmitt, D.; Molenda, C.E.; Hazen, S.L. Myeloperoxidase functions as a major enzymatic catalyst for initiation of lipid peroxidation at sites of inflammation. J. Biol. Chem. 2002, 277, 46116–46122. [Google Scholar] [CrossRef] [Green Version]
  31. Funk, C.D. Prostaglandins and leukotrienes: Advances in eicosanoid biology. Science 2001, 294, 1871–1875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Lands, W.E.; Marshall, P.J.; Kulmacz, R.J. Hydroperoxide availability in the regulation of the arachidonate cascade. Adv. Prostaglandin. Thromboxane. Leukot. Res. 1985, 15, 233–235. [Google Scholar] [PubMed]
  33. Bruegel, M.; Ceglarek, U.; Thiery, J. Eicosanoids: Essential mediators in health and disease/Eicosanoide: Bedeutende Faktoren in der Homöostase und ihre Bedeutung in der Pathogenese multipler Erkrankungen. Lab. Med. 2009, 33, 333–339. [Google Scholar] [CrossRef]
  34. Yui, K.; Imataka, G.; Nakamura, H.; Ohara, N.; Naito, Y. Eicosanoids Derived From Arachidonic Acid and Their Family Prostaglandins and Cyclooxygenase in Psychiatric Disorders. Curr. Neuropharmacol. 2015, 13, 776–785. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Tassoni, D.; Kaur, G.; Weisinger, R.S.; Sinclair, A.J. The role of eicosanoids in the brain. Asia Pac. J. Clin. Nutr. 2008, 17 (Suppl. 1), 220–228. [Google Scholar]
  36. Wymann, M.P.; Schneiter, R. Lipid signalling in disease. Nat. Rev. Mol. Cell Biol. 2008, 9, 162–176. [Google Scholar] [CrossRef]
  37. Ighodaro, O.M.; Akinloye, O.A. First line defence antioxidants-superoxide dismutase (SOD), catalase (CAT) and glutathione peroxidase (GPX): Their fundamental role in the entire antioxidant defence grid. Alex. J. Med. 2018, 54, 287–293. [Google Scholar] [CrossRef] [Green Version]
  38. Birben, E.; Sahiner, U.M.; Sackesen, C.; Erzurum, S.; Kalayci, O. Oxidative stress and antioxidant defense. World Allergy Organ. J. 2012, 5, 9–19. [Google Scholar] [CrossRef] [Green Version]
  39. Liguori, I.; Russo, G.; Curcio, F.; Bulli, G.; Aran, L.; Della-Morte, D.; Gargiulo, G.; Testa, G.; Cacciatore, F.; Bonaduce, D.; et al. Oxidative stress, aging, and diseases. Clin. Interv. Aging 2018, 13, 757–772. [Google Scholar] [CrossRef] [Green Version]
  40. Flatt, T. A new definition of aging? Front. Genet. 2012, 3, 148. [Google Scholar] [CrossRef] [Green Version]
  41. Harman, D. Aging: A theory based on free radical and radiation chemistry. J. Gerontol. 1956, 11, 298–300. [Google Scholar] [CrossRef] [Green Version]
  42. Sanada, F.; Taniyama, Y.; Muratsu, J.; Otsu, R.; Shimizu, H.; Rakugi, H.; Morishita, R. Source of Chronic Inflammation in Aging. Front. Cardiovasc. Med. 2018, 5, 12. [Google Scholar] [CrossRef] [Green Version]
  43. Tan, B.L.; Norhaizan, M.E.; Liew, W.-P.-P.; Sulaiman Rahman, H. Antioxidant and Oxidative Stress: A Mutual Interplay in Age-Related Diseases. Front. Pharmacol. 2018, 9, 1162. [Google Scholar] [CrossRef] [Green Version]
  44. Chung, H.Y.; Kim, D.H.; Lee, E.K.; Chung, K.W.; Chung, S.; Lee, B.; Seo, A.Y.; Chung, J.H.; Jung, Y.S.; Im, E.; et al. Redefining Chronic Inflammation in Aging and Age-Related Diseases: Proposal of the Senoinflammation Concept. Aging Dis. 2019, 10, 367. [Google Scholar] [CrossRef] [Green Version]
  45. Ferrucci, L.; Fabbri, E. Inflammageing: Chronic inflammation in ageing, cardiovascular disease, and frailty. Nat. Rev. Cardiol. 2018, 15, 505–522. [Google Scholar] [CrossRef]
  46. Papaconstantinou, J.; Wang, C.Z.; Zhang, M.; Yang, S.; Deford, J.; Bulavin, D.V.; Ansari, N.H. Attenuation of p38α MAPK stress response signaling delays the in vivo aging of skeletal muscle myofibers and progenitor cells. Aging 2015, 7, 718–733. [Google Scholar] [CrossRef]
  47. Akbari, S.; Amiri, F.T.; Naderi, M.; Shaki, F.; Seyedabadi, M. Sodium arsenite accelerates D-galactose-induced aging in the testis of the rat: Evidence for mitochondrial oxidative damage, NF-kB, JNK, and apoptosis pathways. Toxicology 2022, 470, 153148. [Google Scholar] [CrossRef]
  48. Ng, A.; Tam, W.W.; Zhang, M.W.; Ho, C.S.; Husain, S.F.; McIntyre, R.S.; Ho, R.C. IL-1β, IL-6, TNF- α and CRP in Elderly Patients with Depression or Alzheimer’s disease: Systematic Review and Meta-Analysis. Sci. Rep. 2018, 8, 12050. [Google Scholar] [CrossRef]
  49. Chung, H.Y.; Sung, B.; Jung, K.J.; Zou, Y.; Yu, B.P. The molecular inflammatory process in aging. Antioxid. Redox Signal. 2006, 8, 572–581. [Google Scholar] [CrossRef]
  50. Zhu, Y.; Carvey, P.M.; Ling, Z. Age-related changes in glutathione and glutathione-related enzymes in rat brain. Brain Res. 2006, 1090, 35–44. [Google Scholar] [CrossRef] [Green Version]
  51. Maher, P. The effects of stress and aging on glutathione metabolism. Ageing Res. Rev. 2005, 4, 288–314. [Google Scholar] [CrossRef]
  52. Bellezza, I.; Giambanco, I.; Minelli, A.; Donato, R. Nrf2-Keap1 signaling in oxidative and reductive stress. Biochim. Biophys. Acta Mol. Cell Res. 2018, 1865, 721–733. [Google Scholar] [CrossRef]
  53. Tonelli, C.; Chio, I.I.C.; Tuveson, D.A. Transcriptional Regulation by Nrf2. Antioxid. Redox Signal. 2018, 29, 1727–1745. [Google Scholar] [CrossRef] [Green Version]
  54. Zhang, X.; Guo, J.; Wei, X.; Niu, C.; Jia, M.; Li, Q.; Meng, D. Bach1: Function, Regulation, and Involvement in Disease. Oxid. Med. Cell. Longev. 2018, 2018, 1347969. [Google Scholar] [CrossRef]
  55. Davudian, S.; Mansoori, B.; Shajari, N.; Mohammadi, A.; Baradaran, B. BACH1, the master regulator gene: A novel candidate target for cancer therapy. Gene 2016, 588, 30–37. [Google Scholar] [CrossRef]
  56. Dhakshinamoorthy, S.; Jain, A.K.; Bloom, D.A.; Jaiswal, A.K. Bach1 Competes with Nrf2 Leading to Negative Regulation of the Antioxidant Response Element (ARE)-mediated NAD(P)H:Quinone Oxidoreductase 1 Gene Expression and Induction in Response to Antioxidants *. J. Biol. Chem. 2005, 280, 16891–16900. [Google Scholar] [CrossRef] [Green Version]
  57. Yang, L.; Chen, S.; Zhao, Q.; Sun, Y.; Nie, H. The Critical Role of Bach2 in Shaping the Balance between CD4+ T Cell Subsets in Immune-Mediated Diseases. Mediators Inflamm. 2019, 2019, 2609737. [Google Scholar] [CrossRef] [Green Version]
  58. Tsukumo, S.; Unno, M.; Muto, A.; Takeuchi, A.; Kometani, K.; Kurosaki, T.; Igarashi, K.; Saito, T. Bach2 maintains T cells in a naive state by suppressing effector memory-related genes. Proc. Natl. Acad. Sci. USA 2013, 110, 10735–10740. [Google Scholar] [CrossRef] [Green Version]
  59. Hoshino, H.; Igarashi, K. Expression of the oxidative stress-regulated transcription factor bach2 in differentiating neuronal cells. J. Biochem. 2002, 132, 427–431. [Google Scholar] [CrossRef]
  60. He, X.; Chen, M.G.; Lin, G.X.; Ma, Q. Arsenic induces NAD(P)H-quinone oxidoreductase I by disrupting the Nrf2 x Keap1 x Cul3 complex and recruiting Nrf2 x Maf to the antioxidant response element enhancer. J. Biol. Chem. 2006, 281, 23620–23631. [Google Scholar] [CrossRef] [Green Version]
  61. He, X.; Ma, Q. Critical cysteine residues of Kelch-like ECH-associated protein 1 in arsenic sensing and suppression of nuclear factor erythroid 2-related factor 2. J. Pharmacol. Exp. Ther. 2010, 332, 66–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Yamamoto, T.; Suzuki, T.; Kobayashi, A.; Wakabayashi, J.; Maher, J.; Motohashi, H.; Yamamoto, M. Physiological significance of reactive cysteine residues of Keap1 in determining Nrf2 activity. Mol. Cell. Biol. 2008, 28, 2758–2770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Zhang, D.D.; Hannink, M. Distinct cysteine residues in Keap1 are required for Keap1-dependent ubiquitination of Nrf2 and for stabilization of Nrf2 by chemopreventive agents and oxidative stress. Mol. Cell. Biol. 2003, 23, 8137–8151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Boo, Y.C. Natural Nrf2 Modulators for Skin Protection. Antioxidants 2020, 9, 812. [Google Scholar] [CrossRef]
  65. Osama, A.; Zhang, J.; Yao, J.; Yao, X.; Fang, J. Nrf2: A dark horse in Alzheimer’s disease treatment. Ageing Res. Rev. 2020, 64, 101206. [Google Scholar] [CrossRef]
  66. Yang, S.-Y.; Pyo, M.C.; Nam, M.-H.; Lee, K.-W. ERK/Nrf2 pathway activation by caffeic acid in HepG2 cells alleviates its hepatocellular damage caused by t-butylhydroperoxide-induced oxidative stress. BMC Complement. Altern. Med. 2019, 19, 139. [Google Scholar] [CrossRef] [Green Version]
  67. Hu, Y.-R.; Ma, H.; Zou, Z.-Y.; He, K.; Xiao, Y.-B.; Wang, Y.; Feng, M.; Ye, X.-L.; Li, X.-G. Activation of Akt and JNK/Nrf2/NQO1 pathway contributes to the protective effect of coptisine against AAPH-induced oxidative stress. Biomed. Pharmacother. 2017, 85, 313–322. [Google Scholar] [CrossRef]
  68. Reddy, N.M.; Potteti, H.R.; Vegiraju, S.; Chen, H.-J.; Tamatam, C.M.; Reddy, S.P. PI3K-AKT Signaling via Nrf2 Protects against Hyperoxia-Induced Acute Lung Injury, but Promotes Inflammation Post-Injury Independent of Nrf2 in Mice. PLoS ONE 2015, 10, e0129676. [Google Scholar] [CrossRef] [Green Version]
  69. Lee, O.-H.; Jain, A.K.; Papusha, V.; Jaiswal, A.K. An auto-regulatory loop between stress sensors INrf2 and Nrf2 controls their cellular abundance. J. Biol. Chem. 2007, 282, 36412–36420. [Google Scholar] [CrossRef] [Green Version]
  70. Culbreth, M.; Aschner, M. GSK-3β, a double-edged sword in Nrf2 regulation: Implications for neurological dysfunction and disease. F1000Res 2018, 7, 1043. [Google Scholar] [CrossRef]
  71. Chen, X.; Liu, Y.; Zhu, J.; Lei, S.; Dong, Y.; Li, L.; Jiang, B.; Tan, L.; Wu, J.; Yu, S.; et al. GSK-3β downregulates Nrf2 in cultured cortical neurons and in a rat model of cerebral ischemia-reperfusion. Sci. Rep. 2016, 6, 20196. [Google Scholar] [CrossRef] [Green Version]
  72. Huo, L.; Li, C.-W.; Huang, T.-H.; Lam, Y.C.; Xia, W.; Tu, C.; Chang, W.-C.; Hsu, J.L.; Lee, D.-F.; Nie, L.; et al. Activation of Keap1/Nrf2 signaling pathway by nuclear epidermal growth factor receptor in cancer cells. Am. J. Transl. Res. 2014, 6, 649–663. [Google Scholar]
  73. Tebay, L.E.; Robertson, H.; Durant, S.T.; Vitale, S.R.; Penning, T.M.; Dinkova-Kostova, A.T.; Hayes, J.D. Mechanisms of activation of the transcription factor Nrf2 by redox stressors, nutrient cues, and energy status and the pathways through which it attenuates degenerative disease. Free Radic. Biol. Med. 2015, 88, 108–146. [Google Scholar] [CrossRef] [Green Version]
  74. Fu, J.; Xiong, Z.; Huang, C.; Li, J.; Yang, W.; Han, Y.; Paiboonrungruan, C.; Major, M.B.; Chen, K.-N.; Kang, X.; et al. Hyperactivity of the transcription factor Nrf2 causes metabolic reprogramming in mouse esophagus. J. Biol. Chem. 2019, 294, 327–340. [Google Scholar] [CrossRef] [Green Version]
  75. Zhao, J.; Lin, X.; Meng, D.; Zeng, L.; Zhuang, R.; Huang, S.; Lv, W.; Hu, J. Nrf2 Mediates Metabolic Reprogramming in Non-Small Cell Lung Cancer. Front. Oncol. 2020, 10, 578315. [Google Scholar] [CrossRef]
  76. Kasai, S.; Mimura, J.; Ozaki, T.; Itoh, K. Emerging Regulatory Role of Nrf2 in Iron, Heme, and Hemoglobin Metabolism in Physiology and Disease. Front. Vet. Sci. 2018, 5, 242. [Google Scholar] [CrossRef] [Green Version]
  77. Digaleh, H.; Kiaei, M.; Khodagholi, F. Nrf2 and Nrf1 signaling and ER stress crosstalk: Implication for proteasomal degradation and autophagy. Cell. Mol. Life Sci. 2013, 70, 4681–4694. [Google Scholar] [CrossRef]
  78. Bendavit, G.; Aboulkassim, T.; Hilmi, K.; Shah, S.; Batist, G. Nrf2 Transcription Factor Can Directly Regulate mTOR: Linking Cytoprotective Gene Expression To A Major Metabolic Regulator That Generates Redox Activity. J. Biol. Chem. 2016, 291, 25476–25488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Lau, A.; Wang, X.-J.; Zhao, F.; Villeneuve, N.F.; Wu, T.; Jiang, T.; Sun, Z.; White, E.; Zhang, D.D. A noncanonical mechanism of Nrf2 activation by autophagy deficiency: Direct interaction between Keap1 and p62. Mol. Cell. Biol. 2010, 30, 3275–3285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Jiang, T.; Harder, B.; Rojo de la Vega, M.; Wong, P.K.; Chapman, E.; Zhang, D.D. p62 links autophagy and Nrf2 signaling. Free Radic. Biol. Med. 2015, 88, 199–204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Jain, A.; Lamark, T.; Sjøttem, E.; Larsen, K.B.; Awuh, J.A.; Øvervatn, A.; McMahon, M.; Hayes, J.D.; Johansen, T. p62/SQSTM1 is a target gene for transcription factor NRF2 and creates a positive feedback loop by inducing antioxidant response element-driven gene transcription. J. Biol. Chem. 2010, 285, 22576–22591. [Google Scholar] [CrossRef] [Green Version]
  82. Park, J.S.; Oh, S.Y.; Lee, D.H.; Lee, Y.S.; Sung, S.H.; Ji, H.W.; Lee, M.J.; Lee, Y.-H.; Rhee, S.G.; Bae, S.H. p62/SQSTM1 is required for the protection against endoplasmic reticulum stress-induced apoptotic cell death. Free Radic. Res. 2016, 50, 1408–1421. [Google Scholar] [CrossRef]
  83. Bian, S.; Zhao, Y.; Li, F.; Lu, S.; Yang, S.; Liu, M.; Wang, S.; Zhao, D.; Zhang, W.; Wang, J. Knockdown of p62/sequestosome enhances ginsenoside Rh2-induced apoptosis in cervical cancer HeLa cells with no effect on autophagy. Biosci. Biotechnol. Biochem. 2021, 85, 1097–1103. [Google Scholar] [CrossRef]
  84. Dikovskaya, D.; Dinkova-Kostova, A.T. Measuring Changes in Keap1-Nrf2 Protein Complex Conformation in Individual Cells by FLIM-FRET. Curr. Protoc. Toxicol. 2020, 85, e96. [Google Scholar] [CrossRef]
  85. Dikovskaya, D.; Appleton, P.L.; Bento-Pereira, C.; Dinkova-Kostova, A.T. Measuring the Interaction of Transcription Factor Nrf2 with Its Negative Regulator Keap1 in Single Live Cells by an Improved FRET/FLIM Analysis. Chem. Res. Toxicol. 2019, 32, 500–512. [Google Scholar] [CrossRef]
  86. Baird, L.; Swift, S.; Llères, D.; Dinkova-Kostova, A.T. Monitoring Keap1-Nrf2 interactions in single live cells. Biotechnol. Adv. 2014, 32, 1133–1144. [Google Scholar] [CrossRef] [Green Version]
  87. Heurtaux, T.; Kirchmeyer, M.; Koncina, E.; Felten, P.; Richart, L.; Uriarte Huarte, O.; Schohn, H.; Mittelbronn, M. Apomorphine Reduces A53T α-Synuclein-Induced Microglial Reactivity Through Activation of NRF2 Signalling Pathway. Cell. Mol. Neurobiol. 2021. [Google Scholar] [CrossRef]
  88. Zheng, Y.; Li, L.; Chen, B.; Fang, Y.; Lin, W.; Zhang, T.; Feng, X.; Tao, X.; Wu, Y.; Fu, X.; et al. Chlorogenic acid exerts neuroprotective effect against hypoxia-ischemia brain injury in neonatal rats by activating Sirt1 to regulate the Nrf2-NF-κB signaling pathway. Cell Commun. Signal. 2022, 20, 84. [Google Scholar] [CrossRef]
  89. Zhou, L.; Zhou, M.; Tan, H.; Xiao, M. Cypermethrin-induced cortical neurons apoptosis via the Nrf2/ARE signaling pathway. Pestic. Biochem. Physiol. 2020, 165, 104547. [Google Scholar] [CrossRef]
  90. Zhang, S.; Jin, S.; Zhang, S.; Li, Y.-Y.; Wang, H.; Chen, Y.; Lu, H. Vitexin protects against high glucose-induced endothelial cell apoptosis and oxidative stress via Wnt/β-catenin and Nrf2 signalling pathway. Arch. Physiol. Biochem. 2022, 1–10. [Google Scholar] [CrossRef]
  91. Mutter, F.E.; Park, B.K.; Copple, I.M. Value of monitoring Nrf2 activity for the detection of chemical and oxidative stress. Biochem. Soc. Trans. 2015, 43, 657–662. [Google Scholar] [CrossRef] [Green Version]
  92. Smirnova, N.A.; Haskew-Layton, R.E.; Basso, M.; Hushpulian, D.M.; Payappilly, J.B.; Speer, R.E.; Ahn, Y.-H.; Rakhman, I.; Cole, P.A.; Pinto, J.T.; et al. Development of Neh2-luciferase reporter and its application for high throughput screening and real-time monitoring of Nrf2 activators. Chem. Biol. 2011, 18, 752–765. [Google Scholar] [CrossRef] [Green Version]
  93. Wang, Z.; Chen, G.; Zhu, W.-W.; Zhou, D. Activation of nuclear factor-erythroid 2-related factor 2 (Nrf2) in the basilar artery after subarachnoid hemorrhage in rats. Ann. Clin. Lab. Sci. 2010, 40, 233–239. [Google Scholar]
  94. Zhong, Q.; Mishra, M.; Kowluru, R.A. Transcription factor Nrf2-mediated antioxidant defense system in the development of diabetic retinopathy. Investig. Ophthalmol. Vis. Sci. 2013, 54, 3941–3948. [Google Scholar] [CrossRef]
  95. Wang, X.J.; Hayes, J.D.; Wolf, C.R. Generation of a Stable Antioxidant Response Element–Driven Reporter Gene Cell Line and Its Use to Show Redox-Dependent Activation of Nrf2 by Cancer Chemotherapeutic Agents. Cancer Res. 2006, 66, 10983–10994. [Google Scholar] [CrossRef] [Green Version]
  96. Bischoff, L.J.M.; Kuijper, I.A.; Schimming, J.P.; Wolters, L.; Braak, B.; Langenberg, J.P.; Noort, D.; Beltman, J.B.; van de Water, B. A systematic analysis of Nrf2 pathway activation dynamics during repeated xenobiotic exposure. Arch. Toxicol. 2019, 93, 435–451. [Google Scholar] [CrossRef] [Green Version]
  97. Rahman, I.; Kode, A.; Biswas, S.K. Assay for quantitative determination of glutathione and glutathione disulfide levels using enzymatic recycling method. Nat. Protoc. 2006, 1, 3159–3165. [Google Scholar] [CrossRef]
  98. Wang, J.; Zhu, Q.; Wang, Y.; Peng, J.; Shao, L.; Li, X. Irisin protects against sepsis-associated encephalopathy by suppressing ferroptosis via activation of the Nrf2/GPX4 signal axis. Free Radic. Biol. Med. 2022, 187, 171–184. [Google Scholar] [CrossRef]
  99. Hu, Y.; Wang, P.; Han, K. Hydrogen Attenuated Inflammation Response and Oxidative in Hypoxic Ischemic Encephalopathy via Nrf2 Mediated the Inhibition of NLRP3 and NF-κB. Neuroscience 2022, 485, 23–36. [Google Scholar] [CrossRef]
  100. Oikawa, D.; Akai, R.; Tokuda, M.; Iwawaki, T. A transgenic mouse model for monitoring oxidative stress. Sci. Rep. 2012, 2, 229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Inoue, Y.; Uchiyama, A.; Sekiguchi, A.; Yamazaki, S.; Fujiwara, C.; Yokoyama, Y.; Ogino, S.; Torii, R.; Hosoi, M.; Akai, R.; et al. Protective effect of dimethyl fumarate for the development of pressure ulcers after cutaneous ischemia-reperfusion injury. Wound Repair Regen. 2020, 28, 600–608. [Google Scholar] [CrossRef] [PubMed]
  102. Bazinet, R.P.; Layé, S. Polyunsaturated fatty acids and their metabolites in brain function and disease. Nat. Rev. Neurosci. 2014, 15, 771–785. [Google Scholar] [CrossRef] [PubMed]
  103. Cobley, J.N.; Fiorello, M.L.; Bailey, D.M. 13 reasons why the brain is susceptible to oxidative stress. Redox Biol. 2018, 15, 490–503. [Google Scholar] [CrossRef] [PubMed]
  104. Ren, X.; Zou, L.; Zhang, X.; Branco, V.; Wang, J.; Carvalho, C.; Holmgren, A.; Lu, J. Redox Signaling Mediated by Thioredoxin and Glutathione Systems in the Central Nervous System. Antioxid. Redox Signal. 2017, 27, 989–1010. [Google Scholar] [CrossRef]
  105. Lee, K.H.; Cha, M.; Lee, B.H. Crosstalk between Neuron and Glial Cells in Oxidative Injury and Neuroprotection. Int. J. Mol. Sci. 2021, 22, 13315. [Google Scholar] [CrossRef]
  106. Dringen, R.; Gutterer, J.M.; Hirrlinger, J. Glutathione metabolism in brain: Metabolic interaction between astrocytes and neurons in the defense against reactive oxygen species. Eur. J. Biochem. 2000, 267, 4912–4916. [Google Scholar] [CrossRef]
  107. Chen, Y.; Qin, C.; Huang, J.; Tang, X.; Liu, C.; Huang, K.; Xu, J.; Guo, G.; Tong, A.; Zhou, L. The role of astrocytes in oxidative stress of central nervous system: A mixed blessing. Cell Prolif. 2020, 53, e12781. [Google Scholar] [CrossRef] [Green Version]
  108. Simpson, D.S.A.; Oliver, P.L. ROS Generation in Microglia: Understanding Oxidative Stress and Inflammation in Neurodegenerative Disease. Antioxidants 2020, 9, 743. [Google Scholar] [CrossRef]
  109. Liddell, J.R. Are Astrocytes the Predominant Cell Type for Activation of Nrf2 in Aging and Neurodegeneration? Antioxidants 2017, 6, 65. [Google Scholar] [CrossRef]
  110. He, F.; Ru, X.; Wen, T. NRF2, a Transcription Factor for Stress Response and Beyond. Int. J. Mol. Sci. 2020, 21, 4777. [Google Scholar] [CrossRef]
  111. Boas, S.M.; Joyce, K.L.; Cowell, R.M. The NRF2-Dependent Transcriptional Regulation of Antioxidant Defense Pathways: Relevance for Cell Type-Specific Vulnerability to Neurodegeneration and Therapeutic Intervention. Antioxidants 2021, 11, 8. [Google Scholar] [CrossRef]
  112. Asanuma, M.; Miyazaki, I. Glutathione and Related Molecules in Parkinsonism. Int. J. Mol. Sci. 2021, 22, 8689. [Google Scholar] [CrossRef]
  113. Burton, N.C.; Kensler, T.W.; Guilarte, T.R. In vivo modulation of the Parkinsonian phenotype by Nrf2. Neurotoxicology 2006, 27, 1094–1100. [Google Scholar] [CrossRef]
  114. Hubbs, A.F.; Benkovic, S.A.; Miller, D.B.; O’Callaghan, J.P.; Battelli, L.; Schwegler-Berry, D.; Ma, Q. Vacuolar leukoencephalopathy with widespread astrogliosis in mice lacking transcription factor Nrf2. Am. J. Pathol. 2007, 170, 2068–2076. [Google Scholar] [CrossRef] [Green Version]
  115. Jakel, R.J.; Townsend, J.A.; Kraft, A.D.; Johnson, J.A. Nrf2-mediated protection against 6-hydroxydopamine. Brain Res. 2007, 1144, 192–201. [Google Scholar] [CrossRef] [Green Version]
  116. Innamorato, N.G.; Jazwa, A.; Rojo, A.I.; García, C.; Fernández-Ruiz, J.; Grochot-Przeczek, A.; Stachurska, A.; Jozkowicz, A.; Dulak, J.; Cuadrado, A. Different susceptibility to the Parkinson’s toxin MPTP in mice lacking the redox master regulator Nrf2 or its target gene heme oxygenase-1. PLoS ONE 2010, 5, e11838. [Google Scholar] [CrossRef] [Green Version]
  117. Boyanapalli, S.S.S.; Paredes-Gonzalez, X.; Fuentes, F.; Zhang, C.; Guo, Y.; Pung, D.; Saw, C.L.L.; Kong, A.-N.T. Nrf2 knockout attenuates the anti-inflammatory effects of phenethyl isothiocyanate and curcumin. Chem. Res. Toxicol. 2014, 27, 2036–2043. [Google Scholar] [CrossRef] [Green Version]
  118. Johnson, D.A.; Amirahmadi, S.; Ward, C.; Fabry, Z.; Johnson, J.A. The absence of the pro-antioxidant transcription factor Nrf2 exacerbates experimental autoimmune encephalomyelitis. Toxicol. Sci. 2010, 114, 237–246. [Google Scholar] [CrossRef] [Green Version]
  119. Rojo, A.I.; Pajares, M.; García-Yagüe, A.J.; Buendia, I.; Van Leuven, F.; Yamamoto, M.; López, M.G.; Cuadrado, A. Deficiency in the transcription factor NRF2 worsens inflammatory parameters in a mouse model with combined tauopathy and amyloidopathy. Redox Biol. 2018, 18, 173–180. [Google Scholar] [CrossRef]
  120. Branca, C.; Ferreira, E.; Nguyen, T.-V.; Doyle, K.; Caccamo, A.; Oddo, S. Genetic reduction of Nrf2 exacerbates cognitive deficits in a mouse model of Alzheimer’s disease. Hum. Mol. Genet. 2017, 26, 4823–4835. [Google Scholar] [CrossRef]
  121. Zhu, J.; Wang, H.; Sun, Q.; Ji, X.; Zhu, L.; Cong, Z.; Zhou, Y.; Liu, H.; Zhou, M. Nrf2 is required to maintain the self-renewal of glioma stem cells. BMC Cancer 2013, 13, 380. [Google Scholar] [CrossRef] [Green Version]
  122. Tarantini, S.; Valcarcel-Ares, M.N.; Yabluchanskiy, A.; Tucsek, Z.; Hertelendy, P.; Kiss, T.; Gautam, T.; Zhang, X.A.; Sonntag, W.E.; de Cabo, R.; et al. Nrf2 Deficiency Exacerbates Obesity-Induced Oxidative Stress, Neurovascular Dysfunction, Blood-Brain Barrier Disruption, Neuroinflammation, Amyloidogenic Gene Expression, and Cognitive Decline in Mice, Mimicking the Aging Phenotype. J. Gerontol. A Biol. Sci. Med. Sci. 2018, 73, 853–863. [Google Scholar] [CrossRef]
  123. Sandberg, M.; Patil, J.; D’Angelo, B.; Weber, S.G.; Mallard, C. NRF2-regulation in brain health and disease: Implication of cerebral inflammation. Neuropharmacology 2014, 79, 298–306. [Google Scholar] [CrossRef] [Green Version]
  124. Deshmukh, P.; Unni, S.; Krishnappa, G.; Padmanabhan, B. The Keap1–Nrf2 pathway: Promising therapeutic target to counteract ROS-mediated damage in cancers and neurodegenerative diseases. Biophys. Rev. 2017, 9, 41–56. [Google Scholar] [CrossRef]
  125. Brandes, M.S.; Gray, N.E. NRF2 as a Therapeutic Target in Neurodegenerative Diseases. ASN Neuro 2020, 12, 175909141989978. [Google Scholar] [CrossRef]
  126. Anandhan, A.; Nguyen, N.; Syal, A.; Dreher, L.A.; Dodson, M.; Zhang, D.D.; Madhavan, L. NRF2 Loss Accentuates Parkinsonian Pathology and Behavioral Dysfunction in Human α-Synuclein Overexpressing Mice. Aging Dis. 2021, 12, 964–982. [Google Scholar] [CrossRef]
  127. Bono, S.; Feligioni, M.; Corbo, M. Impaired antioxidant KEAP1-NRF2 system in amyotrophic lateral sclerosis: NRF2 activation as a potential therapeutic strategy. Mol. Neurodegener. 2021, 16, 71. [Google Scholar] [CrossRef]
  128. Cuadrado, A. Brain-Protective Mechanisms of Transcription Factor NRF2: Toward a Common Strategy for Neurodegenerative Diseases. Annu. Rev. Pharmacol. Toxicol. 2022, 62, 255–277. [Google Scholar] [CrossRef]
  129. Furman, D.; Campisi, J.; Verdin, E.; Carrera-Bastos, P.; Targ, S.; Franceschi, C.; Ferrucci, L.; Gilroy, D.W.; Fasano, A.; Miller, G.W.; et al. Chronic inflammation in the etiology of disease across the life span. Nat. Med. 2019, 25, 1822–1832. [Google Scholar] [CrossRef]
  130. Liu, T.; Zhang, L.; Joo, D.; Sun, S.-C. NF-κB signaling in inflammation. Signal Transduct. Target. Ther. 2017, 2, 17023. [Google Scholar] [CrossRef] [Green Version]
  131. Christian, F.; Smith, E.L.; Carmody, R.J. The Regulation of NF-κB Subunits by Phosphorylation. Cells 2016, 5, 12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Wardyn, J.D.; Ponsford, A.H.; Sanderson, C.M. Dissecting molecular cross-talk between Nrf2 and NF-κB response pathways. Biochem. Soc. Trans. 2015, 43, 621–626. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Sivandzade, F.; Prasad, S.; Bhalerao, A.; Cucullo, L. NRF2 and NF-κB interplay in cerebrovascular and neurodegenerative disorders: Molecular mechanisms and possible therapeutic approaches. Redox Biol. 2019, 21, 101059. [Google Scholar] [CrossRef] [PubMed]
  134. Ganesh Yerra, V.; Negi, G.; Sharma, S.S.; Kumar, A. Potential therapeutic effects of the simultaneous targeting of the Nrf2 and NF-κB pathways in diabetic neuropathy. Redox Biol. 2013, 1, 394–397. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Li, W.; Khor, T.O.; Xu, C.; Shen, G.; Jeong, W.-S.; Yu, S.; Kong, A.-N. Activation of Nrf2-antioxidant signaling attenuates NFkappaB-inflammatory response and elicits apoptosis. Biochem. Pharmacol. 2008, 76, 1485–1489. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Ahmed, S.M.U.; Luo, L.; Namani, A.; Wang, X.J.; Tang, X. Nrf2 signaling pathway: Pivotal roles in inflammation. Biochim. Biophys. Acta Mol. Basis Dis. 2017, 1863, 585–597. [Google Scholar] [CrossRef] [PubMed]
  137. Saha, S.; Buttari, B.; Panieri, E.; Profumo, E.; Saso, L. An Overview of Nrf2 Signaling Pathway and Its Role in Inflammation. Molecules 2020, 25, 5474. [Google Scholar] [CrossRef]
  138. Pae, H.-O.; Chung, H.-T. Heme oxygenase-1: Its therapeutic roles in inflammatory diseases. Immune Netw. 2009, 9, 12–19. [Google Scholar] [CrossRef] [Green Version]
  139. Drummond, G.S.; Baum, J.; Greenberg, M.; Lewis, D.; Abraham, N.G. HO-1 overexpression and underexpression: Clinical implications. Arch. Biochem. Biophys. 2019, 673, 108073. [Google Scholar] [CrossRef]
  140. Subedi, L.; Lee, J.H.; Yumnam, S.; Ji, E.; Kim, S.Y. Anti-Inflammatory Effect of Sulforaphane on LPS-Activated Microglia Potentially through JNK/AP-1/NF-κB Inhibition and Nrf2/HO-1 Activation. Cells 2019, 8, 194. [Google Scholar] [CrossRef] [Green Version]
  141. Nitti, M.; Furfaro, A.L.; Mann, G.E. Heme Oxygenase Dependent Bilirubin Generation in Vascular Cells: A Role in Preventing Endothelial Dysfunction in Local Tissue Microenvironment? Front. Physiol. 2020, 11, 23. [Google Scholar] [CrossRef]
  142. Helou, D.G.; Martin, S.F.; Pallardy, M.; Chollet-Martin, S.; Kerdine-Römer, S. Nrf2 Involvement in Chemical-Induced Skin Innate Immunity. Front. Immunol. 2019, 10, 1004. [Google Scholar] [CrossRef] [Green Version]
  143. Krajka-Kuźniak, V.; Baer-Dubowska, W. Modulation of Nrf2 and NF-κB Signaling Pathways by Naturally Occurring Compounds in Relation to Cancer Prevention and Therapy. Are Combinations Better Than Single Compounds? Int. J. Mol. Sci. 2021, 22, 8223. [Google Scholar] [CrossRef]
  144. Yu, M.; Li, H.; Liu, Q.; Liu, F.; Tang, L.; Li, C.; Yuan, Y.; Zhan, Y.; Xu, W.; Li, W.; et al. Nuclear factor p65 interacts with Keap1 to repress the Nrf2-ARE pathway. Cell. Signal. 2011, 23, 883–892. [Google Scholar] [CrossRef]
  145. Hu, B.; Wei, H.; Song, Y.; Chen, M.; Fan, Z.; Qiu, R.; Zhu, W.; Xu, W.; Wang, F. NF-κB and Keap1 Interaction Represses Nrf2-Mediated Antioxidant Response in Rabbit Hemorrhagic Disease Virus Infection. J. Virol. 2020, 94, e00016-20. [Google Scholar] [CrossRef]
  146. Bellezza, I.; Mierla, A.L.; Minelli, A. Nrf2 and NF-κB and Their Concerted Modulation in Cancer Pathogenesis and Progression. Cancers 2010, 2, 483–497. [Google Scholar] [CrossRef]
  147. Guo, X.; Hong, S.; He, H.; Zeng, Y.; Chen, Y.; Mo, X.; Li, J.; Li, L.; Steinmetz, R.; Liu, Q. NFκB promotes oxidative stress-induced necrosis and ischemia/reperfusion injury by inhibiting Nrf2-ARE pathway. Free Radic. Biol. Med. 2020, 159, 125–135. [Google Scholar] [CrossRef]
  148. Gao, W.; Guo, L.; Yang, Y.; Wang, Y.; Xia, S.; Gong, H.; Zhang, B.-K.; Yan, M. Dissecting the Crosstalk Between Nrf2 and NF-κB Response Pathways in Drug-Induced Toxicity. Front. Cell Dev. Biol. 2021, 9, 809952. [Google Scholar] [CrossRef]
  149. Liu, G.-H.; Qu, J.; Shen, X. NF-kappaB/p65 antagonizes Nrf2-ARE pathway by depriving CBP from Nrf2 and facilitating recruitment of HDAC3 to MafK. Biochim. Biophys. Acta 2008, 1783, 713–727. [Google Scholar] [CrossRef] [Green Version]
  150. Jin, W.; Zhu, L.; Guan, Q.; Chen, G.; Wang, Q.F.; Yin, H.X.; Hang, C.H.; Shi, J.X.; Wang, H.D. Influence of Nrf2 genotype on pulmonary NF-kappaB activity and inflammatory response after traumatic brain injury. Ann. Clin. Lab. Sci. 2008, 38, 221–227. [Google Scholar]
  151. Pan, H.; Wang, H.; Wang, X.; Zhu, L.; Mao, L. The absence of Nrf2 enhances NF-κB-dependent inflammation following scratch injury in mouse primary cultured astrocytes. Mediators Inflamm. 2012, 2012, 217580. [Google Scholar] [CrossRef]
  152. Wang, K.; Zheng, M.; Lester, K.L.; Han, Z. Light-induced Nrf2-/- mice as atrophic age-related macular degeneration model and treatment with nanoceria laden injectable hydrogel. Sci. Rep. 2019, 9, 14573. [Google Scholar] [CrossRef]
  153. Chang, S.-H.; Lee, J.-S.; Yun, U.J.; Park, K.W. A Role of Stress Sensor Nrf2 in Stimulating Thermogenesis and Energy Expenditure. Biomedicines 2021, 9, 1196. [Google Scholar] [CrossRef]
  154. Jardim, N.S.; Müller, S.G.; Pase, F.M.; Nogueira, C.W. Nuclear Factor [Erythroid-derived 2]-like 2 and Mitochondrial Transcription Factor A Contribute to Moderate-intensity Swimming Effectiveness against Memory Impairment in Young Mice Induced by Concomitant Exposure to a High-calorie Diet during the Early Life Period. Neuroscience 2021, 452, 311–325. [Google Scholar] [CrossRef] [PubMed]
  155. Stevenson, R.J.; Francis, H.M. The hippocampus and the regulation of human food intake. Psychol. Bull. 2017, 143, 1011–1032. [Google Scholar] [CrossRef]
  156. Mattson, M.P. Energy intake and exercise as determinants of brain health and vulnerability to injury and disease. Cell Metab. 2012, 16, 706–722. [Google Scholar] [CrossRef] [Green Version]
  157. Vasconcelos, A.R.; Dos Santos, N.B.; Scavone, C.; Munhoz, C.D. Nrf2/ARE Pathway Modulation by Dietary Energy Regulation in Neurological Disorders. Front. Pharmacol. 2019, 10, 33. [Google Scholar] [CrossRef] [Green Version]
  158. Matsui, S.; Sasaki, T.; Kohno, D.; Yaku, K.; Inutsuka, A.; Yokota-Hashimoto, H.; Kikuchi, O.; Suga, T.; Kobayashi, M.; Yamanaka, A.; et al. Neuronal SIRT1 regulates macronutrient-based diet selection through FGF21 and oxytocin signalling in mice. Nat. Commun. 2018, 9, 4604. [Google Scholar] [CrossRef] [Green Version]
  159. Gälman, C.; Lundåsen, T.; Kharitonenkov, A.; Bina, H.A.; Eriksson, M.; Hafström, I.; Dahlin, M.; Amark, P.; Angelin, B.; Rudling, M. The circulating metabolic regulator FGF21 is induced by prolonged fasting and PPARalpha activation in man. Cell Metab. 2008, 8, 169–174. [Google Scholar] [CrossRef] [Green Version]
  160. Laeger, T.; Henagan, T.M.; Albarado, D.C.; Redman, L.M.; Bray, G.A.; Noland, R.C.; Münzberg, H.; Hutson, S.M.; Gettys, T.W.; Schwartz, M.W.; et al. FGF21 is an endocrine signal of protein restriction. J. Clin. Investig. 2014, 124, 3913–3922. [Google Scholar] [CrossRef] [Green Version]
  161. Solon-Biet, S.M.; Cogger, V.C.; Pulpitel, T.; Heblinski, M.; Wahl, D.; McMahon, A.C.; Warren, A.; Durrant-Whyte, J.; Walters, K.A.; Krycer, J.R.; et al. Defining the Nutritional and Metabolic Context of FGF21 Using the Geometric Framework. Cell Metab. 2016, 24, 555–565. [Google Scholar] [CrossRef] [Green Version]
  162. Lundsgaard, A.-M.; Fritzen, A.M.; Sjøberg, K.A.; Myrmel, L.S.; Madsen, L.; Wojtaszewski, J.F.P.; Richter, E.A.; Kiens, B. Circulating FGF21 in humans is potently induced by short term overfeeding of carbohydrates. Mol. Metab. 2017, 6, 22–29. [Google Scholar] [CrossRef] [PubMed]
  163. Jensen-Cody, S.O.; Flippo, K.H.; Claflin, K.E.; Yavuz, Y.; Sapouckey, S.A.; Walters, G.C.; Usachev, Y.M.; Atasoy, D.; Gillum, M.P.; Potthoff, M.J. FGF21 Signals to Glutamatergic Neurons in the Ventromedial Hypothalamus to Suppress Carbohydrate Intake. Cell Metab. 2020, 32, 273–286.e6. [Google Scholar] [CrossRef] [PubMed]
  164. Sajja, R.K.; Prasad, S.; Tang, S.; Kaisar, M.A.; Cucullo, L. Blood-brain barrier disruption in diabetic mice is linked to Nrf2 signaling deficits: Role of ABCB10? Neurosci. Lett. 2017, 653, 152–158. [Google Scholar] [CrossRef]
  165. Yu, Z.; Lin, L.; Jiang, Y.; Chin, I.; Wang, X.; Li, X.; Lo, E.H.; Wang, X. Recombinant FGF21 Protects Against Blood-Brain Barrier Leakage Through Nrf2 Upregulation in Type 2 Diabetes Mice. Mol. Neurobiol. 2019, 56, 2314–2327. [Google Scholar] [CrossRef]
  166. Furusawa, Y.; Uruno, A.; Yagishita, Y.; Higashi, C.; Yamamoto, M. Nrf2 induces fibroblast growth factor 21 in diabetic mice. Genes Cells 2014, 19, 864–878. [Google Scholar] [CrossRef]
  167. Sáenz de Urturi, D.; Buqué, X.; Porteiro, B.; Folgueira, C.; Mora, A.; Delgado, T.C.; Prieto-Fernández, E.; Olaizola, P.; Gómez-Santos, B.; Apodaka-Biguri, M.; et al. Methionine adenosyltransferase 1a antisense oligonucleotides activate the liver-brown adipose tissue axis preventing obesity and associated hepatosteatosis. Nat. Commun. 2022, 13, 1096. [Google Scholar] [CrossRef]
  168. Yagishita, Y.; Uruno, A.; Fukutomi, T.; Saito, R.; Saigusa, D.; Pi, J.; Fukamizu, A.; Sugiyama, F.; Takahashi, S.; Yamamoto, M. Nrf2 Improves Leptin and Insulin Resistance Provoked by Hypothalamic Oxidative Stress. Cell Rep. 2017, 18, 2030–2044. [Google Scholar] [CrossRef] [Green Version]
  169. Bösl, M.R.; Seldin, M.F.; Nishimura, S.; Taketo, M. Cloning, structural analysis and mapping of the mouse selenocysteine tRNA([Ser]Sec) gene (Trsp). Mol. Gen. Genet. 1995, 248, 247–252. [Google Scholar] [CrossRef]
  170. Shen, G.; Zhou, L.; Liu, W.; Cui, Y.; Xie, W.; Chen, H.; Yu, W.; Li, W.; Li, H. Di(2-ethylhexyl)phthalate Alters the Synthesis and β-Oxidation of Fatty Acids and Hinders ATP Supply in Mouse Testes via UPLC-Q-Exactive Orbitrap MS-Based Metabonomics Study. J. Agric. Food Chem. 2017, 65, 5056–5063. [Google Scholar] [CrossRef]
  171. Wang, Y.; Li, L.; Wang, Y.; Zhu, X.; Jiang, M.; Song, E.; Song, Y. New application of the commercial sweetener rebaudioside a as a hepatoprotective candidate: Induction of the Nrf2 signaling pathway. Eur. J. Pharmacol. 2018, 822, 128–137. [Google Scholar] [CrossRef] [PubMed]
  172. Xia, S.-F.; Shao, J.; Zhao, S.-Y.; Qiu, Y.-Y.; Teng, L.-P.; Huang, W.; Wang, S.-S.; Cheng, X.-R.; Jiang, Y.-Y. Niga-ichigoside F1 ameliorates high-fat diet-induced hepatic steatosis in male mice by Nrf2 activation. Food Funct. 2018, 9, 906–916. [Google Scholar] [CrossRef] [PubMed]
  173. Niu, Y.; He, J.; Ahmad, H.; Shen, M.; Zhao, Y.; Gan, Z.; Zhang, L.; Zhong, X.; Wang, C.; Wang, T. Dietary Curcumin Supplementation Increases Antioxidant Capacity, Upregulates Nrf2 and Hmox1 Levels in the Liver of Piglet Model with Intrauterine Growth Retardation. Nutrients 2019, 11, 2978. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Amos, D.; Cook, C.; Santanam, N. Omega 3 rich diet modulates energy metabolism via GPR120-Nrf2 crosstalk in a novel antioxidant mouse model. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2019, 1864, 466–488. [Google Scholar] [CrossRef]
  175. Rajendran, P.; Ammar, R.B.; Al-Saeedi, F.J.; Mohamed, M.E.; ElNaggar, M.A.; Al-Ramadan, S.Y.; Bekhet, G.M.; Soliman, A.M. Kaempferol Inhibits Zearalenone-Induced Oxidative Stress and Apoptosis via the PI3K/Akt-Mediated Nrf2 Signaling Pathway: In Vitro and In Vivo Studies. Int. J. Mol. Sci. 2020, 22, 217. [Google Scholar] [CrossRef]
  176. Erten, F.; Orhan, C.; Tuzcu, M.; Er, B.; Defo Deeh, P.B.; Sahin, N.; Özercan, I.H.; Juturu, V.; Sahin, K. Salacia chinensis exerts its antidiabetic effect by modulating glucose-regulated proteins and transcription factors in high-fat diet fed-streptozotocin-induced type 2 diabetic rats. J. Food Biochem. 2020, 44, e13513. [Google Scholar] [CrossRef]
  177. Illesca, P.; Valenzuela, R.; Espinosa, A.; Echeverría, F.; Soto-Alarcon, S.; Campos, C.; Rodriguez, A.; Vargas, R.; Magrone, T.; Videla, L.A. Protective Effects of Eicosapentaenoic Acid Plus Hydroxytyrosol Supplementation Against White Adipose Tissue Abnormalities in Mice Fed a High-Fat Diet. Molecules 2020, 25, 4433. [Google Scholar] [CrossRef]
  178. Abrescia, P.; Treppiccione, L.; Rossi, M.; Bergamo, P. Modulatory role of dietary polyunsaturated fatty acids in Nrf2-mediated redox homeostasis. Prog. Lipid Res. 2020, 80, 101066. [Google Scholar] [CrossRef]
  179. Mohamed, A.B.; Rémond, D.; Gual-Grau, A.; Bernalier-Donnadille, A.; Capel, F.; Michalski, M.-C.; Laugerette, F.; Cohade, B.; Hafnaoui, N.; Béchet, D.; et al. A Mix of Dietary Fibres Changes Interorgan Nutrients Exchanges and Muscle-Adipose Energy Handling in Overfed Mini-Pigs. Nutrients 2021, 13, 4202. [Google Scholar] [CrossRef]
  180. Bruns, D.R.; Drake, J.C.; Biela, L.M.; Peelor, F.F.; Miller, B.F.; Hamilton, K.L. Nrf2 Signaling and the Slowed Aging Phenotype: Evidence from Long-Lived Models. Oxid. Med. Cell. Longev. 2015, 2015, 732596. [Google Scholar] [CrossRef] [Green Version]
  181. Dodson, M.; Anandhan, A.; Zhang, D.D.; Madhavan, L. An NRF2 Perspective on Stem Cells and Ageing. Front. Aging 2021, 2, 690686. [Google Scholar] [CrossRef]
  182. Matsumaru, D.; Motohashi, H. The KEAP1-NRF2 System in Healthy Aging and Longevity. Antioxidants 2021, 10, 1929. [Google Scholar] [CrossRef]
  183. Lewis, K.N.; Wason, E.; Edrey, Y.H.; Kristan, D.M.; Nevo, E.; Buffenstein, R. Regulation of Nrf2 signaling and longevity in naturally long-lived rodents. Proc. Natl. Acad. Sci. USA 2015, 112, 3722–3727. [Google Scholar] [CrossRef] [Green Version]
  184. Zhang, H.; Davies, K.J.A.; Forman, H.J. Oxidative stress response and Nrf2 signaling in aging. Free Radic. Biol. Med. 2015, 88, 314–336. [Google Scholar] [CrossRef] [Green Version]
  185. Silva-Palacios, A.; Ostolga-Chavarría, M.; Zazueta, C.; Königsberg, M. Nrf2: Molecular and epigenetic regulation during aging. Ageing Res. Rev. 2018, 47, 31–40. [Google Scholar] [CrossRef]
  186. Tsay, H.J.; Wang, P.; Wang, S.L.; Ku, H.H. Age-associated changes of superoxide dismutase and catalase activities in the rat brain. J. Biomed. Sci. 2000, 7, 466–474. [Google Scholar] [CrossRef]
  187. Samarghandian, S.; Azimi-Nezhad, M.; Samini, F. Preventive effect of safranal against oxidative damage in aged male rat brain. Exp. Anim. 2015, 64, 65–71. [Google Scholar] [CrossRef] [Green Version]
  188. Berr, C.; Balansard, B.; Arnaud, J.; Roussel, A.M.; Alpérovitch, A. Cognitive decline is associated with systemic oxidative stress: The EVA study. Etude du Vieillissement Artériel. J. Am. Geriatr. Soc. 2000, 48, 1285–1291. [Google Scholar] [CrossRef]
  189. Baierle, M.; Nascimento, S.N.; Moro, A.M.; Brucker, N.; Freitas, F.; Gauer, B.; Durgante, J.; Bordignon, S.; Zibetti, M.; Trentini, C.M.; et al. Relationship between inflammation and oxidative stress and cognitive decline in the institutionalized elderly. Oxid. Med. Cell. Longev. 2015, 2015, 804198. [Google Scholar] [CrossRef] [Green Version]
  190. Schöttker, B.; Saum, K.-U.; Jansen, E.H.J.M.; Holleczek, B.; Brenner, H. Associations of metabolic, inflammatory and oxidative stress markers with total morbidity and multi-morbidity in a large cohort of older German adults. Age Ageing 2016, 45, 127–135. [Google Scholar] [CrossRef] [Green Version]
  191. Hajjar, I.; Hayek, S.S.; Goldstein, F.C.; Martin, G.; Jones, D.P.; Quyyumi, A. Oxidative stress predicts cognitive decline with aging in healthy adults: An observational study. J. Neuroinflamm. 2018, 15, 17. [Google Scholar] [CrossRef] [Green Version]
  192. Cuadrado, A.; Manda, G.; Hassan, A.; Alcaraz, M.J.; Barbas, C.; Daiber, A.; Ghezzi, P.; León, R.; López, M.G.; Oliva, B.; et al. Transcription Factor NRF2 as a Therapeutic Target for Chronic Diseases: A Systems Medicine Approach. Pharmacol. Rev. 2018, 70, 348–383. [Google Scholar] [CrossRef] [Green Version]
  193. Hushpulian, D.M.; Ammal Kaidery, N.; Ahuja, M.; Poloznikov, A.A.; Sharma, S.M.; Gazaryan, I.G.; Thomas, B. Challenges and Limitations of Targeting the Keap1-Nrf2 Pathway for Neurotherapeutics: Bach1 De-Repression to the Rescue. Front. Aging Neurosci. 2021, 13, 673205. [Google Scholar] [CrossRef]
  194. Han, K.; Jin, X.; Guo, X.; Cao, G.; Tian, S.; Song, Y.; Zuo, Y.; Yu, P.; Gao, G.; Chang, Y.-Z. Nrf2 knockout altered brain iron deposition and mitigated age-related motor dysfunction in aging mice. Free Radic. Biol. Med. 2021, 162, 592–602. [Google Scholar] [CrossRef]
  195. Liu, Z.; Han, K.; Huo, X.; Yan, B.; Gao, M.; Lv, X.; Yu, P.; Gao, G.; Chang, Y.-Z. Nrf2 knockout dysregulates iron metabolism and increases the hemolysis through ROS in aging mice. Life Sci. 2020, 255, 117838. [Google Scholar] [CrossRef]
  196. Bao, W.-D.; Pang, P.; Zhou, X.-T.; Hu, F.; Xiong, W.; Chen, K.; Wang, J.; Wang, F.; Xie, D.; Hu, Y.-Z.; et al. Loss of ferroportin induces memory impairment by promoting ferroptosis in Alzheimer’s disease. Cell Death Differ. 2021, 28, 1548–1562. [Google Scholar] [CrossRef] [PubMed]
  197. Raha, A.A.; Biswas, A.; Henderson, J.; Chakraborty, S.; Holland, A.; Friedland, R.P.; Mukaetova-Ladinska, E.; Zaman, S.; Raha-Chowdhury, R. Interplay of Ferritin Accumulation and Ferroportin Loss in Ageing Brain: Implication for Protein Aggregation in Down Syndrome Dementia, Alzheimer’s, and Parkinson’s Diseases. Int. J. Mol. Sci. 2022, 23, 1060. [Google Scholar] [CrossRef] [PubMed]
  198. Kubben, N.; Zhang, W.; Wang, L.; Voss, T.C.; Yang, J.; Qu, J.; Liu, G.-H.; Misteli, T. Repression of the Antioxidant NRF2 Pathway in Premature Aging. Cell 2016, 165, 1361–1374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Sadowska-Bartosz, I.; Bartosz, G. Effect of antioxidants supplementation on aging and longevity. BioMed Res. Int. 2014, 2014, 404680. [Google Scholar] [CrossRef] [PubMed]
  200. Conti, V.; Izzo, V.; Corbi, G.; Russomanno, G.; Manzo, V.; De Lise, F.; Di Donato, A.; Filippelli, A. Antioxidant Supplementation in the Treatment of Aging-Associated Diseases. Front. Pharmacol. 2016, 7, 24. [Google Scholar] [CrossRef] [Green Version]
  201. Grilc, N.K.; Sova, M.; Kristl, J. Drug Delivery Strategies for Curcumin and Other Natural Nrf2 Modulators of Oxidative Stress-Related Diseases. Pharmaceutics 2021, 13, 2137. [Google Scholar] [CrossRef]
  202. Malavolta, M.; Bracci, M.; Santarelli, L.; Sayeed, M.A.; Pierpaoli, E.; Giacconi, R.; Costarelli, L.; Piacenza, F.; Basso, A.; Cardelli, M.; et al. Inducers of Senescence, Toxic Compounds, and Senolytics: The Multiple Faces of Nrf2-Activating Phytochemicals in Cancer Adjuvant Therapy. Mediators Inflamm. 2018, 2018, 4159013. [Google Scholar] [CrossRef] [Green Version]
  203. Kulkarni, A.S.; Gubbi, S.; Barzilai, N. Benefits of Metformin in Attenuating the Hallmarks of Aging. Cell Metab. 2020, 32, 15–30. [Google Scholar] [CrossRef]
  204. Ngoi, N.Y.L.; Liew, A.Q.X.; Chong, S.J.F.; Davids, M.S.; Clement, M.-V.; Pervaiz, S. The redox-senescence axis and its therapeutic targeting. Redox Biol. 2021, 45, 102032. [Google Scholar] [CrossRef]
  205. Zhang, L.; Pitcher, L.E.; Prahalad, V.; Niedernhofer, L.J.; Robbins, P.D. Targeting cellular senescence with senotherapeutics: Senolytics and senomorphics. FEBS J. 2022, 197, 111526. [Google Scholar] [CrossRef]
  206. Nayeri Rad, A.; Shams, G.; Avelar, R.A.; Morowvat, M.H.; Ghasemi, Y. Potential senotherapeutic candidates and their combinations derived from transcriptional connectivity and network measures. Inform. Med. Unlocked 2022, 30, 100920. [Google Scholar] [CrossRef]
  207. Sikora, E.; Bielak-Zmijewska, A.; Dudkowska, M.; Krzystyniak, A.; Mosieniak, G.; Wesierska, M.; Wlodarczyk, J. Cellular Senescence in Brain Aging. Front. Aging Neurosci. 2021, 13, 646924. [Google Scholar] [CrossRef]
  208. Martínez-Cué, C.; Rueda, N. Cellular Senescence in Neurodegenerative Diseases. Front. Cell. Neurosci. 2020, 14, 16. [Google Scholar] [CrossRef]
  209. Ogrodnik, M.; Evans, S.A.; Fielder, E.; Victorelli, S.; Kruger, P.; Salmonowicz, H.; Weigand, B.M.; Patel, A.D.; Pirtskhalava, T.; Inman, C.L.; et al. Whole-body senescent cell clearance alleviates age-related brain inflammation and cognitive impairment in mice. Aging Cell 2021, 20, e13296. [Google Scholar] [CrossRef]
  210. Krzystyniak, A.; Wesierska, M.; Petrazzo, G.; Gadecka, A.; Dudkowska, M.; Bielak-Zmijewska, A.; Mosieniak, G.; Figiel, I.; Wlodarczyk, J.; Sikora, E. Combination of dasatinib and quercetin improves cognitive abilities in aged male Wistar rats, alleviates inflammation and changes hippocampal synaptic plasticity and histone H3 methylation profile. Aging 2022, 14, 572–595. [Google Scholar] [CrossRef]
  211. Yuan, H.; Xu, Y.; Luo, Y.; Wang, N.-X.; Xiao, J.-H. Role of Nrf2 in cell senescence regulation. Mol. Cell. Biochem. 2021, 476, 247–259. [Google Scholar] [CrossRef]
  212. Saha, S.; Buttari, B.; Profumo, E.; Tucci, P.; Saso, L. A Perspective on Nrf2 Signaling Pathway for Neuroinflammation: A Potential Therapeutic Target in Alzheimer’s and Parkinson’s Diseases. Front. Cell. Neurosci. 2022, 15, 787258. [Google Scholar] [CrossRef]
  213. Uddin, M.S.; Stachowiak, A.; Mamun, A.A.; Tzvetkov, N.T.; Takeda, S.; Atanasov, A.G.; Bergantin, L.B.; Abdel-Daim, M.M.; Stankiewicz, A.M. Autophagy and Alzheimer’s Disease: From Molecular Mechanisms to Therapeutic Implications. Front. Aging Neurosci. 2018, 10, 4. [Google Scholar] [CrossRef]
  214. Dinkova-Kostova, A.T.; Kostov, R.V.; Kazantsev, A.G. The role of Nrf2 signaling in counteracting neurodegenerative diseases. FEBS J. 2018, 285, 3576–3590. [Google Scholar] [CrossRef] [Green Version]
  215. Schmidlin, C.J.; Dodson, M.B.; Madhavan, L.; Zhang, D.D. Redox Regulation by NRF2 in Aging and Disease. Free Radic. Biol. Med. 2019, 134, 702–707. [Google Scholar] [CrossRef]
  216. Joshi, G.; Gan, K.A.; Johnson, D.A.; Johnson, J.A. Increased Alzheimer’s disease-like pathology in the APP/ PS1ΔE9 mouse model lacking Nrf2 through modulation of autophagy. Neurobiol. Aging 2015, 36, 664–679. [Google Scholar] [CrossRef] [Green Version]
  217. Ren, P.; Chen, J.; Li, B.; Zhang, M.; Yang, B.; Guo, X.; Chen, Z.; Cheng, H.; Wang, P.; Wang, S.; et al. Nrf2 Ablation Promotes Alzheimer’s Disease-Like Pathology in APP/PS1 Transgenic Mice: The Role of Neuroinflammation and Oxidative Stress. Oxid. Med. Cell. Longev. 2020, 2020, 3050971. [Google Scholar] [CrossRef]
  218. Kanninen, K.; Heikkinen, R.; Malm, T.; Rolova, T.; Kuhmonen, S.; Leinonen, H.; Ylä-Herttuala, S.; Tanila, H.; Levonen, A.-L.; Koistinaho, M.; et al. Intrahippocampal injection of a lentiviral vector expressing Nrf2 improves spatial learning in a mouse model of Alzheimer’s disease. Proc. Natl. Acad. Sci. USA 2009, 106, 16505–16510. [Google Scholar] [CrossRef] [Green Version]
  219. Jiwaji, Z.; Tiwari, S.S.; Avilés-Reyes, R.X.; Hooley, M.; Hampton, D.; Torvell, M.; Johnson, D.A.; McQueen, J.; Baxter, P.; Sabari-Sankar, K.; et al. Reactive astrocytes acquire neuroprotective as well as deleterious signatures in response to Tau and Aß pathology. Nat. Commun. 2022, 13, 135. [Google Scholar] [CrossRef]
  220. Ramsey, C.P.; Glass, C.A.; Montgomery, M.B.; Lindl, K.A.; Ritson, G.P.; Chia, L.A.; Hamilton, R.L.; Chu, C.T.; Jordan-Sciutto, K.L. Expression of Nrf2 in neurodegenerative diseases. J. Neuropathol. Exp. Neurol. 2007, 66, 75–85. [Google Scholar] [CrossRef]
  221. Delaidelli, A.; Richner, M.; Jiang, L.; van der Laan, A.; Bergholdt Jul Christiansen, I.; Ferreira, N.; Nyengaard, J.R.; Vægter, C.B.; Jensen, P.H.; Mackenzie, I.R.; et al. α-Synuclein pathology in Parkinson disease activates homeostatic NRF2 anti-oxidant response. Acta Neuropathol. Commun. 2021, 9, 105. [Google Scholar] [CrossRef] [PubMed]
  222. Barone, M.C.; Sykiotis, G.P.; Bohmann, D. Genetic activation of Nrf2 signaling is sufficient to ameliorate neurodegenerative phenotypes in a Drosophila model of Parkinson’s disease. Dis. Model. Mech. 2011, 4, 701–707. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Zhang, C.; Zhao, M.; Wang, B.; Su, Z.; Guo, B.; Qin, L.; Zhang, W.; Zheng, R. The Nrf2-NLRP3-caspase-1 axis mediates the neuroprotective effects of Celastrol in Parkinson’s disease. Redox Biol. 2021, 47, 102134. [Google Scholar] [CrossRef] [PubMed]
  224. Zhong, Y.; Cai, X.; Ding, L.; Liao, J.; Liu, X.; Huang, Y.; Chen, X.; Long, L. Nrf2 Inhibits the Progression of Parkinson’s Disease by Upregulating AABR07032261.5 to Repress Pyroptosis. J. Inflamm. Res. 2022, 15, 669–685. [Google Scholar] [CrossRef] [PubMed]
  225. Chen, J.; Rusnak, M.; Lombroso, P.J.; Sidhu, A. Dopamine promotes striatal neuronal apoptotic death via ERK signaling cascades. Eur. J. Neurosci. 2009, 29, 287–306. [Google Scholar] [CrossRef] [Green Version]
  226. Rojo, A.I.; Innamorato, N.G.; Martín-Moreno, A.M.; De Ceballos, M.L.; Yamamoto, M.; Cuadrado, A. Nrf2 regulates microglial dynamics and neuroinflammation in experimental Parkinson’s disease. Glia 2010, 58, 588–598. [Google Scholar] [CrossRef]
  227. Lastres-Becker, I.; Ulusoy, A.; Innamorato, N.G.; Sahin, G.; Rábano, A.; Kirik, D.; Cuadrado, A. α-Synuclein expression and Nrf2 deficiency cooperate to aggravate protein aggregation, neuronal death and inflammation in early-stage Parkinson’s disease. Hum. Mol. Genet. 2012, 21, 3173–3192. [Google Scholar] [CrossRef] [Green Version]
  228. Gan, L.; Vargas, M.R.; Johnson, D.A.; Johnson, J.A. Astrocyte-specific overexpression of Nrf2 delays motor pathology and synuclein aggregation throughout the CNS in the alpha-synuclein mutant (A53T) mouse model. J. Neurosci. 2012, 32, 17775–17787. [Google Scholar] [CrossRef] [Green Version]
  229. Vargas, M.R.; Burton, N.C.; Kutzke, J.; Gan, L.; Johnson, D.A.; Schäfer, M.; Werner, S.; Johnson, J.A. Absence of Nrf2 or its selective overexpression in neurons and muscle does not affect survival in ALS-linked mutant hSOD1 mouse models. PLoS ONE 2013, 8, e56625. [Google Scholar] [CrossRef]
  230. Sigfridsson, E.; Marangoni, M.; Johnson, J.A.; Hardingham, G.E.; Fowler, J.H.; Horsburgh, K. Astrocyte-specific overexpression of Nrf2 protects against optic tract damage and behavioural alterations in a mouse model of cerebral hypoperfusion. Sci. Rep. 2018, 8, 12552. [Google Scholar] [CrossRef] [Green Version]
  231. Daniels, C.M.L.; Austin, E.V.; Rockney, D.E.; Jacka, E.M.; Hagemann, T.L.; Johnson, D.A.; Johnson, J.A.; Messing, A. Beneficial Effects of Nrf2 Overexpression in a Mouse Model of Alexander Disease. J. Neurosci. 2012, 32, 10507–10515. [Google Scholar] [CrossRef] [Green Version]
  232. Jenner, P. Oxidative stress in Parkinson’s disease. Ann. Neurol. 2003, 53 (Suppl. 3), S26–S36; discussion S36–S38. [Google Scholar] [CrossRef]
  233. Gan, L.; Johnson, J.A. Oxidative damage and the Nrf2-ARE pathway in neurodegenerative diseases. Biochim. Biophys. Acta 2014, 1842, 1208–1218. [Google Scholar] [CrossRef] [Green Version]
  234. Vinish, M.; Anand, A.; Prabhakar, S. Altered oxidative stress levels in Indian Parkinson’s disease patients with PARK2 mutations. Acta Biochim. Pol. 2011, 58, 165–169. [Google Scholar] [CrossRef] [Green Version]
  235. Manoharan, S.; Guillemin, G.J.; Abiramasundari, R.S.; Essa, M.M.; Akbar, M.; Akbar, M.D. The Role of Reactive Oxygen Species in the Pathogenesis of Alzheimer’s Disease, Parkinson’s Disease, and Huntington’s Disease: A Mini Review. Oxid. Med. Cell. Longev. 2016, 2016, 8590578. [Google Scholar] [CrossRef]
  236. Wei, Z.; Li, X.; Li, X.; Liu, Q.; Cheng, Y. Oxidative Stress in Parkinson’s Disease: A Systematic Review and Meta-Analysis. Front. Mol. Neurosci. 2018, 11, 236. [Google Scholar] [CrossRef]
  237. Pan, H.; Wang, H.; Zhu, L.; Wang, X.; Cong, Z.; Sun, K.; Fan, Y. The involvement of Nrf2-ARE pathway in regulation of apoptosis in human glioblastoma cell U251. Neurol. Res. 2013, 35, 71–78. [Google Scholar] [CrossRef]
  238. Zhou, Y.; Wang, H.-D.; Zhu, L.; Cong, Z.-X.; Li, N.; Ji, X.-J.; Pan, H.; Wang, J.-W.; Li, W.-C. Knockdown of Nrf2 enhances autophagy induced by temozolomide in U251 human glioma cell line. Oncol. Rep. 2013, 29, 394–400. [Google Scholar] [CrossRef] [Green Version]
  239. Cong, Z.-X.; Wang, H.-D.; Wang, J.-W.; Zhou, Y.; Pan, H.; Zhang, D.-D.; Zhu, L. ERK and PI3K signaling cascades induce Nrf2 activation and regulate cell viability partly through Nrf2 in human glioblastoma cells. Oncol. Rep. 2013, 30, 715–722. [Google Scholar] [CrossRef] [Green Version]
  240. Fan, Z.; Wirth, A.-K.; Chen, D.; Wruck, C.J.; Rauh, M.; Buchfelder, M.; Savaskan, N. Nrf2-Keap1 pathway promotes cell proliferation and diminishes ferroptosis. Oncogenesis 2017, 6, e371. [Google Scholar] [CrossRef] [Green Version]
  241. Ji, X.-J.; Chen, S.-H.; Zhu, L.; Pan, H.; Zhou, Y.; Li, W.; You, W.-C.; Gao, C.-C.; Zhu, J.-H.; Jiang, K.; et al. Knockdown of NF-E2-related factor 2 inhibits the proliferation and growth of U251MG human glioma cells in a mouse xenograft model. Oncol. Rep. 2013, 30, 157–164. [Google Scholar] [CrossRef] [PubMed]
  242. Ji, X.; Wang, H.; Zhu, J.; Zhu, L.; Pan, H.; Li, W.; Zhou, Y.; Cong, Z.; Yan, F.; Chen, S. Knockdown of Nrf2 suppresses glioblastoma angiogenesis by inhibiting hypoxia-induced activation of HIF-1α. Int. J. Cancer 2014, 135, 574–584. [Google Scholar] [CrossRef] [PubMed]
  243. Haapasalo, J.; Nordfors, K.; Granberg, K.J.; Kivioja, T.; Nykter, M.; Haapasalo, H.; Soini, Y. NRF2, DJ1 and SNRX1 and their prognostic impact in astrocytic gliomas. Histol. Histopathol. 2018, 33, 791–801. [Google Scholar] [CrossRef] [PubMed]
  244. Ji, X.; Wang, H.; Zhu, J.; Tang, Y.; Zhou, Y.; Zhu, L.; Gao, C.; Li, W.; You, W.; Yu, B.; et al. Correlation of Nrf2 and HIF-1α in glioblastoma and their relationships to clinicopathologic features and survival. Neurol. Res. 2013, 35, 1044–1050. [Google Scholar] [CrossRef]
  245. Pölönen, P.; Jawahar Deen, A.; Leinonen, H.M.; Jyrkkänen, H.-K.; Kuosmanen, S.; Mononen, M.; Jain, A.; Tuomainen, T.; Pasonen-Seppänen, S.; Hartikainen, J.M.; et al. Nrf2 and SQSTM1/p62 jointly contribute to mesenchymal transition and invasion in glioblastoma. Oncogene 2019, 38, 7473–7490. [Google Scholar] [CrossRef] [Green Version]
  246. Kanamori, M.; Higa, T.; Sonoda, Y.; Murakami, S.; Dodo, M.; Kitamura, H.; Taguchi, K.; Shibata, T.; Watanabe, M.; Suzuki, H.; et al. Activation of the NRF2 pathway and its impact on the prognosis of anaplastic glioma patients. Neuro-Oncology 2015, 17, 555–565. [Google Scholar] [CrossRef] [Green Version]
  247. Li, K.; Ouyang, L.; He, M.; Luo, M.; Cai, W.; Tu, Y.; Pi, R.; Liu, A. IDH1 R132H mutation regulates glioma chemosensitivity through Nrf2 pathway. Oncotarget 2017, 8, 28865–28879. [Google Scholar] [CrossRef] [Green Version]
  248. Ahmad, F.; Dixit, D.; Sharma, V.; Kumar, A.; Joshi, S.D.; Sarkar, C.; Sen, E. Nrf2-driven TERT regulates pentose phosphate pathway in glioblastoma. Cell Death Dis. 2016, 7, e2213. [Google Scholar] [CrossRef] [Green Version]
  249. Tsakonas, G.; Martín-Bernabé, A.; Rounis, K.; Moreno-Ruiz, P.; Botling, J.; De Petris, L.; Ylipää, A.; Mezheyeuski, A.; Micke, P.; Östman, A.; et al. High Density of NRF2 Expression in Malignant Cells Is Associated with Increased Risk of CNS Metastasis in Early-Stage NSCLC. Cancers 2021, 13, 3151. [Google Scholar] [CrossRef]
  250. Aljohani, H.M.; Aittaleb, M.; Furgason, J.M.; Amaya, P.; Deeb, A.; Chalmers, J.J.; Bahassi, E.M. Genetic mutations associated with lung cancer metastasis to the brain. Mutagenesis 2018, 33, 137–145. [Google Scholar] [CrossRef]
  251. Cong, Z.-X.; Wang, H.-D.; Zhou, Y.; Wang, J.-W.; Pan, H.; Zhang, D.-D.; Zhang, L.; Zhu, L. Temozolomide and irradiation combined treatment-induced Nrf2 activation increases chemoradiation sensitivity in human glioblastoma cells. J. Neurooncol. 2014, 116, 41–48. [Google Scholar] [CrossRef]
  252. Rocha, C.R.R.; Reily Rocha, A.; Molina Silva, M.; Rodrigues Gomes, L.; Teatin Latancia, M.; Andrade Tomaz, M.; de Souza, I.; Karolynne Seregni Monteiro, L.; Menck, C.F.M. Revealing Temozolomide Resistance Mechanisms via Genome-Wide CRISPR Libraries. Cells 2020, 9, 2573. [Google Scholar] [CrossRef]
  253. Escoll, M.; Lastra, D.; Robledinos-Antón, N.; Wandosell, F.; Antón, I.M.; Cuadrado, A. WIP Modulates Oxidative Stress through NRF2/KEAP1 in Glioblastoma Cells. Antioxidants 2020, 9, 773. [Google Scholar] [CrossRef]
  254. Beltzig, L.; Schwarzenbach, C.; Leukel, P.; Frauenknecht, K.B.M.; Sommer, C.; Tancredi, A.; Hegi, M.E.; Christmann, M.; Kaina, B. Senescence Is the Main Trait Induced by Temozolomide in Glioblastoma Cells. Cancers 2022, 14, 2233. [Google Scholar] [CrossRef]
  255. Yagishita, Y.; Gatbonton-Schwager, T.N.; McCallum, M.L.; Kensler, T.W. Current Landscape of NRF2 Biomarkers in Clinical Trials. Antioxidants 2020, 9, 716. [Google Scholar] [CrossRef]
  256. Kornberg, R.D.; Lorch, Y. Twenty-five years of the nucleosome, fundamental particle of the eukaryote chromosome. Cell 1999, 98, 285–294. [Google Scholar] [CrossRef] [Green Version]
  257. Baker, M. Making sense of chromatin states. Nat. Methods 2011, 8, 717–722. [Google Scholar] [CrossRef]
  258. Ohnmacht, J.; May, P.; Sinkkonen, L.; Krüger, R. Missing heritability in Parkinson’s disease: The emerging role of non-coding genetic variation. J. Neural Transm. 2020, 127, 729–748. [Google Scholar] [CrossRef] [Green Version]
  259. Blackwood, E.M.; Kadonaga, J.T. Going the distance: A current view of enhancer action. Science 1998, 281, 60–63. [Google Scholar] [CrossRef] [Green Version]
  260. Krivega, I.; Dean, A. Enhancer and promoter interactions-long distance calls. Curr. Opin. Genet. Dev. 2012, 22, 79–85. [Google Scholar] [CrossRef] [Green Version]
  261. Correa, F.; Mallard, C.; Nilsson, M.; Sandberg, M. Activated microglia decrease histone acetylation and Nrf2-inducible anti-oxidant defence in astrocytes: Restoring effects of inhibitors of HDACs, p38 MAPK and GSK3β. Neurobiol. Dis. 2011, 44, 142–151. [Google Scholar] [CrossRef] [Green Version]
  262. Wang, B.; Zhu, X.; Kim, Y.; Li, J.; Huang, S.; Saleem, S.; Li, R.; Xu, Y.; Dore, S.; Cao, W. Histone deacetylase inhibition activates transcription factor Nrf2 and protects against cerebral ischemic damage. Free Radic. Biol. Med. 2012, 52, 928–936. [Google Scholar] [CrossRef]
  263. Li, Z.; Xu, L.; Tang, N.; Xu, Y.; Ye, X.; Shen, S.; Niu, X.; Lu, S.; Chen, Z. The polycomb group protein EZH2 inhibits lung cancer cell growth by repressing the transcription factor Nrf2. FEBS Lett. 2014, 588, 3000–3007. [Google Scholar] [CrossRef] [Green Version]
  264. Chorley, B.N.; Campbell, M.R.; Wang, X.; Karaca, M.; Sambandan, D.; Bangura, F.; Xue, P.; Pi, J.; Kleeberger, S.R.; Bell, D.A. Identification of novel NRF2-regulated genes by ChIP-Seq: Influence on retinoid X receptor alpha. Nucleic Acids Res. 2012, 40, 7416–7429. [Google Scholar] [CrossRef] [Green Version]
  265. Malhotra, D.; Portales-Casamar, E.; Singh, A.; Srivastava, S.; Arenillas, D.; Happel, C.; Shyr, C.; Wakabayashi, N.; Kensler, T.W.; Wasserman, W.W.; et al. Global mapping of binding sites for Nrf2 identifies novel targets in cell survival response through ChIP-Seq profiling and network analysis. Nucleic Acids Res. 2010, 38, 5718–5734. [Google Scholar] [CrossRef]
  266. He, C.H.; Gong, P.; Hu, B.; Stewart, D.; Choi, M.E.; Choi, A.M.; Alam, J. Identification of activating transcription factor 4 (ATF4) as an Nrf2-interacting protein. Implication for heme oxygenase-1 gene regulation. J. Biol. Chem. 2001, 276, 20858–20865. [Google Scholar] [CrossRef] [Green Version]
  267. Zhang, J.; Ohta, T.; Maruyama, A.; Hosoya, T.; Nishikawa, K.; Maher, J.M.; Shibahara, S.; Itoh, K.; Yamamoto, M. BRG1 interacts with Nrf2 to selectively mediate HO-1 induction in response to oxidative stress. Mol. Cell. Biol. 2006, 26, 7942–7952. [Google Scholar] [CrossRef] [Green Version]
  268. Alam, J.; Killeen, E.; Gong, P.; Naquin, R.; Hu, B.; Stewart, D.; Ingelfinger, J.R.; Nath, K.A. Heme activates the heme oxygenase-1 gene in renal epithelial cells by stabilizing Nrf2. Am. J. Physiol. Renal Physiol. 2003, 284, F743–F752. [Google Scholar] [CrossRef] [Green Version]
  269. Sekine, H.; Okazaki, K.; Ota, N.; Shima, H.; Katoh, Y.; Suzuki, N.; Igarashi, K.; Ito, M.; Motohashi, H.; Yamamoto, M. The Mediator Subunit MED16 Transduces NRF2-Activating Signals into Antioxidant Gene Expression. Mol. Cell. Biol. 2016, 36, 407–420. [Google Scholar] [CrossRef] [Green Version]
  270. Nioi, P.; Nguyen, T.; Sherratt, P.J.; Pickett, C.B. The carboxy-terminal Neh3 domain of Nrf2 is required for transcriptional activation. Mol. Cell. Biol. 2005, 25, 10895–10906. [Google Scholar] [CrossRef] [Green Version]
  271. Kim, J.-H.; Yu, S.; Chen, J.D.; Kong, A.N. The nuclear cofactor RAC3/AIB1/SRC-3 enhances Nrf2 signaling by interacting with transactivation domains. Oncogene 2013, 32, 514–527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  272. Pan, H.; Guan, D.; Liu, X.; Li, J.; Wang, L.; Wu, J.; Zhou, J.; Zhang, W.; Ren, R.; Zhang, W.; et al. SIRT6 safeguards human mesenchymal stem cells from oxidative stress by coactivating NRF2. Cell Res. 2016, 26, 190–205. [Google Scholar] [CrossRef] [PubMed]
  273. Guttman, M.; Amit, I.; Garber, M.; French, C.; Lin, M.F.; Feldser, D.; Huarte, M.; Zuk, O.; Carey, B.W.; Cassady, J.P.; et al. Chromatin signature reveals over a thousand highly conserved large non-coding RNAs in mammals. Nature 2009, 458, 223–227. [Google Scholar] [CrossRef] [PubMed]
  274. Ikeda, Y.; Sugawara, A.; Taniyama, Y.; Uruno, A.; Igarashi, K.; Arima, S.; Ito, S.; Takeuchi, K. Suppression of rat thromboxane synthase gene transcription by peroxisome proliferator-activated receptor gamma in macrophages via an interaction with NRF2. J. Biol. Chem. 2000, 275, 33142–33150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  275. Shah, N.M.; Rushworth, S.A.; Murray, M.Y.; Bowles, K.M.; MacEwan, D.J. Understanding the role of NRF2-regulated miRNAs in human malignancies. Oncotarget 2013, 4, 1130–1142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Joo, M.S.; Lee, C.G.; Koo, J.H.; Kim, S.G. miR-125b transcriptionally increased by Nrf2 inhibits AhR repressor, which protects kidney from cisplatin-induced injury. Cell Death Dis. 2013, 4, e899. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  277. Kwak, M.-K.; Wakabayashi, N.; Itoh, K.; Motohashi, H.; Yamamoto, M.; Kensler, T.W. Modulation of gene expression by cancer chemopreventive dithiolethiones through the Keap1-Nrf2 pathway. Identification of novel gene clusters for cell survival. J. Biol. Chem. 2003, 278, 8135–8145. [Google Scholar] [CrossRef] [Green Version]
  278. Handy, D.E.; Castro, R.; Loscalzo, J. Epigenetic modifications: Basic mechanisms and role in cardiovascular disease. Circulation 2011, 123, 2145–2156. [Google Scholar] [CrossRef] [Green Version]
  279. Marstrand, T.T.; Storey, J.D. Identifying and mapping cell-type-specific chromatin programming of gene expression. Proc. Natl. Acad. Sci. USA 2014, 111, E645–E654. [Google Scholar] [CrossRef] [Green Version]
  280. Winick-Ng, W.; Kukalev, A.; Harabula, I.; Zea-Redondo, L.; Szabó, D.; Meijer, M.; Serebreni, L.; Zhang, Y.; Bianco, S.; Chiariello, A.M.; et al. Cell-type specialization is encoded by specific chromatin topologies. Nature 2021, 599, 684–691. [Google Scholar] [CrossRef]
  281. Czamara, D.; Eraslan, G.; Page, C.M.; Lahti, J.; Lahti-Pulkkinen, M.; Hämäläinen, E.; Kajantie, E.; Laivuori, H.; Villa, P.M.; Reynolds, R.M.; et al. Integrated analysis of environmental and genetic influences on cord blood DNA methylation in new-borns. Nat. Commun. 2019, 10, 2548. [Google Scholar] [CrossRef] [Green Version]
  282. Breton, C.V.; Landon, R.; Kahn, L.G.; Enlow, M.B.; Peterson, A.K.; Bastain, T.; Braun, J.; Comstock, S.S.; Duarte, C.S.; Hipwell, A.; et al. Exploring the evidence for epigenetic regulation of environmental influences on child health across generations. Commun. Biol. 2021, 4, 769. [Google Scholar] [CrossRef]
  283. Wang, Z.; Pan, Q.; Gendron, P.; Zhu, W.; Guo, F.; Cen, S.; Wainberg, M.A.; Liang, C. CRISPR/Cas9-Derived Mutations Both Inhibit HIV-1 Replication and Accelerate Viral Escape. Cell Rep. 2016, 15, 481–489. [Google Scholar] [CrossRef] [Green Version]
  284. Levings, D.C.; Wang, X.; Kohlhase, D.; Bell, D.A.; Slattery, M. A distinct class of antioxidant response elements is consistently activated in tumors with NRF2 mutations. Redox Biol. 2018, 19, 235–249. [Google Scholar] [CrossRef]
  285. Oyake, T.; Itoh, K.; Motohashi, H.; Hayashi, N.; Hoshino, H.; Nishizawa, M.; Yamamoto, M.; Igarashi, K. Bach proteins belong to a novel family of BTB-basic leucine zipper transcription factors that interact with MafK and regulate transcription through the NF-E2 site. Mol. Cell. Biol. 1996, 16, 6083–6095. [Google Scholar] [CrossRef] [Green Version]
  286. Itoh-Nakadai, A.; Hikota, R.; Muto, A.; Kometani, K.; Watanabe-Matsui, M.; Sato, Y.; Kobayashi, M.; Nakamura, A.; Miura, Y.; Yano, Y.; et al. The transcription repressors Bach2 and Bach1 promote B cell development by repressing the myeloid program. Nat. Immunol. 2014, 15, 1171–1180. [Google Scholar] [CrossRef]
  287. Shin, S.; Wakabayashi, N.; Misra, V.; Biswal, S.; Lee, G.H.; Agoston, E.S.; Yamamoto, M.; Kensler, T.W. NRF2 modulates aryl hydrocarbon receptor signaling: Influence on adipogenesis. Mol. Cell. Biol. 2007, 27, 7188–7197. [Google Scholar] [CrossRef] [Green Version]
  288. Chan, K.; Lu, R.; Chang, J.C.; Kan, Y.W. NRF2, a member of the NFE2 family of transcription factors, is not essential for murine erythropoiesis, growth, and development. Proc. Natl. Acad. Sci. USA 1996, 93, 13943–13948. [Google Scholar] [CrossRef] [Green Version]
  289. Rushworth, S.A.; Zaitseva, L.; Murray, M.Y.; Shah, N.M.; Bowles, K.M.; MacEwan, D.J. The high Nrf2 expression in human acute myeloid leukemia is driven by NF-κB and underlies its chemo-resistance. Blood 2012, 120, 5188–5198. [Google Scholar] [CrossRef] [Green Version]
  290. Tao, S.; Wang, S.; Moghaddam, S.J.; Ooi, A.; Chapman, E.; Wong, P.K.; Zhang, D.D. Oncogenic KRAS confers chemoresistance by upregulating NRF2. Cancer Res. 2014, 74, 7430–7441. [Google Scholar] [CrossRef] [Green Version]
  291. Saunders, A.; Macosko, E.Z.; Wysoker, A.; Goldman, M.; Krienen, F.M.; de Rivera, H.; Bien, E.; Baum, M.; Bortolin, L.; Wang, S.; et al. Molecular Diversity and Specializations among the Cells of the Adult Mouse Brain. Cell 2018, 174, 1015–1030.e16. [Google Scholar] [CrossRef] [Green Version]
  292. Murphy, T.H.; Yu, J.; Ng, R.; Johnson, D.A.; Shen, H.; Honey, C.R.; Johnson, J.A. Preferential expression of antioxidant response element mediated gene expression in astrocytes. J. Neurochem. 2001, 76, 1670–1678. [Google Scholar] [CrossRef]
  293. Shih, A.Y.; Johnson, D.A.; Wong, G.; Kraft, A.D.; Jiang, L.; Erb, H.; Johnson, J.A.; Murphy, T.H. Coordinate Regulation of Glutathione Biosynthesis and Release by Nrf2-Expressing Glia Potently Protects Neurons from Oxidative Stress. J. Neurosci. 2003, 23, 3394–3406. [Google Scholar] [CrossRef]
  294. Kraft, A.D.; Johnson, D.A.; Johnson, J.A. Nuclear Factor E2-Related Factor 2-Dependent Antioxidant Response Element Activation by tert-Butylhydroquinone and Sulforaphane Occurring Preferentially in Astrocytes Conditions Neurons against Oxidative Insult. J. Neurosci. 2004, 24, 1101–1112. [Google Scholar] [CrossRef] [Green Version]
  295. Bell, K.F.S.; Al-Mubarak, B.; Martel, M.-A.; McKay, S.; Wheelan, N.; Hasel, P.; Márkus, N.M.; Baxter, P.; Deighton, R.F.; Serio, A.; et al. Neuronal development is promoted by weakened intrinsic antioxidant defences due to epigenetic repression of Nrf2. Nat. Commun. 2015, 6, 7066. [Google Scholar] [CrossRef] [Green Version]
  296. Narasimhan, M.; Patel, D.; Vedpathak, D.; Rathinam, M.; Henderson, G.; Mahimainathan, L. Identification of novel microRNAs in post-transcriptional control of Nrf2 expression and redox homeostasis in neuronal, SH-SY5Y cells. PLoS ONE 2012, 7, e51111. [Google Scholar] [CrossRef] [Green Version]
  297. Jing, X.; Shi, H.; Zhang, C.; Ren, M.; Han, M.; Wei, X.; Zhang, X.; Lou, H. Dimethyl fumarate attenuates 6-OHDA-induced neurotoxicity in SH-SY5Y cells and in animal model of Parkinson’s disease by enhancing Nrf2 activity. Neuroscience 2015, 286, 131–140. [Google Scholar] [CrossRef]
  298. Ahuja, M.; Ammal Kaidery, N.; Yang, L.; Calingasan, N.; Smirnova, N.; Gaisin, A.; Gaisina, I.N.; Gazaryan, I.; Hushpulian, D.M.; Kaddour-Djebbar, I.; et al. Distinct Nrf2 Signaling Mechanisms of Fumaric Acid Esters and Their Role in Neuroprotection against 1-Methyl-4-Phenyl-1,2,3,6-Tetrahydropyridine-Induced Experimental Parkinson’s-Like Disease. J. Neurosci. 2016, 36, 6332–6351. [Google Scholar] [CrossRef] [Green Version]
  299. Campolo, M.; Casili, G.; Biundo, F.; Crupi, R.; Cordaro, M.; Cuzzocrea, S.; Esposito, E. The Neuroprotective Effect of Dimethyl Fumarate in an MPTP-Mouse Model of Parkinson’s Disease: Involvement of Reactive Oxygen Species/Nuclear Factor-κB/Nuclear Transcription Factor Related to NF-E2. Antioxid. Redox Signal. 2017, 27, 453–471. [Google Scholar] [CrossRef] [Green Version]
  300. Campolo, M.; Casili, G.; Lanza, M.; Filippone, A.; Paterniti, I.; Cuzzocrea, S.; Esposito, E. Multiple mechanisms of dimethyl fumarate in amyloid β-induced neurotoxicity in human neuronal cells. J. Cell. Mol. Med. 2018, 22, 1081–1094. [Google Scholar] [CrossRef]
  301. Rajput, M.S.; Nirmal, N.P.; Rathore, D.; Dahima, R. Dimethyl Fumarate Mitigates Tauopathy in Aβ-Induced Neuroblastoma SH-SY5Y Cells. Neurochem. Res. 2020, 45, 2641–2652. [Google Scholar] [CrossRef] [PubMed]
  302. Sun, X.; Suo, X.; Xia, X.; Yu, C.; Dou, Y. Dimethyl Fumarate is a Potential Therapeutic Option for Alzheimer’s Disease. J. Alzheimers Dis. 2022, 85, 443–456. [Google Scholar] [CrossRef] [PubMed]
  303. Singh, A.; Venkannagari, S.; Oh, K.H.; Zhang, Y.-Q.; Rohde, J.M.; Liu, L.; Nimmagadda, S.; Sudini, K.; Brimacombe, K.R.; Gajghate, S.; et al. Small Molecule Inhibitor of NRF2 Selectively Intervenes Therapeutic Resistance in KEAP1-Deficient NSCLC Tumors. ACS Chem. Biol. 2016, 11, 3214–3225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Okazaki, K.; Anzawa, H.; Liu, Z.; Ota, N.; Kitamura, H.; Onodera, Y.; Alam, M.M.; Matsumaru, D.; Suzuki, T.; Katsuoka, F.; et al. Enhancer remodeling promotes tumor-initiating activity in NRF2-activated non-small cell lung cancers. Nat. Commun. 2020, 11, 5911. [Google Scholar] [CrossRef]
  305. Preissl, S.; Fang, R.; Huang, H.; Zhao, Y.; Raviram, R.; Gorkin, D.U.; Zhang, Y.; Sos, B.C.; Afzal, V.; Dickel, D.E.; et al. Single-nucleus analysis of accessible chromatin in developing mouse forebrain reveals cell-type-specific transcriptional regulation. Nat. Neurosci. 2018, 21, 432–439. [Google Scholar] [CrossRef] [Green Version]
  306. Mich, J.K.; Graybuck, L.T.; Hess, E.E.; Mahoney, J.T.; Kojima, Y.; Ding, Y.; Somasundaram, S.; Miller, J.A.; Kalmbach, B.E.; Radaelli, C.; et al. Functional enhancer elements drive subclass-selective expression from mouse to primate neocortex. Cell Rep. 2021, 34, 108754. [Google Scholar] [CrossRef]
  307. Gui, Y.; Grzyb, K.; Thomas, M.H.; Ohnmacht, J.; Garcia, P.; Buttini, M.; Skupin, A.; Sauter, T.; Sinkkonen, L. Single-nuclei chromatin profiling of ventral midbrain reveals cell identity transcription factors and cell-type-specific gene regulatory variation. Epigenetics Chromatin 2021, 14, 43. [Google Scholar] [CrossRef]
  308. Ahuja, M.; Ammal Kaidery, N.; Attucks, O.C.; McDade, E.; Hushpulian, D.M.; Gaisin, A.; Gaisina, I.; Ahn, Y.H.; Nikulin, S.; Poloznikov, A.; et al. Bach1 derepression is neuroprotective in a mouse model of Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2021, 118, e2111643118. [Google Scholar] [CrossRef]
  309. Wang, X.J.; Hayes, J.D.; Henderson, C.J.; Wolf, C.R. Identification of retinoic acid as an inhibitor of transcription factor Nrf2 through activation of retinoic acid receptor alpha. Proc. Natl. Acad. Sci. USA 2007, 104, 19589–19594. [Google Scholar] [CrossRef] [Green Version]
  310. Armand, E.J.; Li, J.; Xie, F.; Luo, C.; Mukamel, E.A. Single-Cell Sequencing of Brain Cell Transcriptomes and Epigenomes. Neuron 2021, 109, 11–26. [Google Scholar] [CrossRef]
  311. Maurano, M.T.; Humbert, R.; Rynes, E.; Thurman, R.E.; Haugen, E.; Wang, H.; Reynolds, A.P.; Sandstrom, R.; Qu, H.; Brody, J.; et al. Systematic localization of common disease-associated variation in regulatory DNA. Science 2012, 337, 1190–1195. [Google Scholar] [CrossRef] [Green Version]
  312. Schaub, M.A.; Boyle, A.P.; Kundaje, A.; Batzoglou, S.; Snyder, M. Linking disease associations with regulatory information in the human genome. Genome Res. 2012, 22, 1748–1759. [Google Scholar] [CrossRef] [Green Version]
  313. Deplancke, B.; Alpern, D.; Gardeux, V. The Genetics of Transcription Factor DNA Binding Variation. Cell 2016, 166, 538–554. [Google Scholar] [CrossRef] [Green Version]
  314. Rojo de la Vega, M.; Chapman, E.; Zhang, D.D. NRF2 and the Hallmarks of Cancer. Cancer Cell 2018, 34, 21–43. [Google Scholar] [CrossRef]
  315. Tsai, W.-C.; Hueng, D.-Y.; Lin, C.-R.; Yang, T.C.K.; Gao, H.-W. Nrf2 Expressions Correlate with WHO Grades in Gliomas and Meningiomas. Int. J. Mol. Sci. 2016, 17, 722. [Google Scholar] [CrossRef] [Green Version]
  316. Wu, S.; Lu, H.; Bai, Y. Nrf2 in cancers: A double-edged sword. Cancer Med. 2019, 8, 2252–2267. [Google Scholar] [CrossRef]
Figure 1. Reactive species production and disruption of the redox homeostasis. (A) Generation of superoxide anions (O2) by the complexes I and III of the mitochondrial respiratory chain. (B) ROS are byproducts of the normal metabolism of oxygen. (1) ROS formation follows the monoelectronic reduction of O2. (2) Superoxide dismutases catalyse the dismutation of superoxide into hydrogen peroxide H2O2. (3) Hydroxyl radical HO is produced by decomposition of H2O2 via the Fenton reaction (Fe(II)-dependent reaction). (4) Glutathione peroxidase catalyses the reduction of hydrogen peroxide to water via oxidation of reduced glutathione (GSH) into its disulphide form (GSSG). (5) Catalase reacts with the hydrogen peroxide to catalyse the formation of water and O2. (6) The production of nitric oxide (NO), a reactive nitrogen species, is carried out from L-arginine by nitric oxide synthases. (7) Peroxynitrite ONOO is produced by the reaction of the free radical superoxide O2●− with the free radical nitric oxide NO. (8) Hypochlorous acid (HOCl) is produced through myeloperoxidase (MPO)-catalysed peroxidation of chloride anions using H2O2. (9) Nitrogen dioxide (NO2) reacts with molecular oxygen to form, under the action of UV radiation and heat, nitric oxide (NO) and ozone (O3). Lipid peroxidation (10)–(12) is a chain of reactions of oxidative degradation of lipids: (10) unsaturated lipid (LH) reacts with prooxidants (hydroxyl radical, HO), leading to the formation of lipid radical (L) and water; (11) lipid radical then reacts with oxygen to form a lipid peroxy radical (LOO), which abstracts one hydrogen from another lipid molecule, generating a new lipid radical (L) and lipid hydroperoxide (LOOH) (12). (C) The redox balance is an equilibrium between ROS production, due to pro-oxidant conditions, and antioxidant defences. A disruption of this redox balance in favour of oxidative conditions will promote an oxidative stress. CAT, catalase; CoQ, coenzyme Q; Cyt C, cytochrome C; GPx, glutathione peroxidase; GR, glutathione reductase; GSH, reduced glutathione; GSSG, oxidized glutathione; SOD, superoxide dismutases; Trx, thioredoxin.
Figure 1. Reactive species production and disruption of the redox homeostasis. (A) Generation of superoxide anions (O2) by the complexes I and III of the mitochondrial respiratory chain. (B) ROS are byproducts of the normal metabolism of oxygen. (1) ROS formation follows the monoelectronic reduction of O2. (2) Superoxide dismutases catalyse the dismutation of superoxide into hydrogen peroxide H2O2. (3) Hydroxyl radical HO is produced by decomposition of H2O2 via the Fenton reaction (Fe(II)-dependent reaction). (4) Glutathione peroxidase catalyses the reduction of hydrogen peroxide to water via oxidation of reduced glutathione (GSH) into its disulphide form (GSSG). (5) Catalase reacts with the hydrogen peroxide to catalyse the formation of water and O2. (6) The production of nitric oxide (NO), a reactive nitrogen species, is carried out from L-arginine by nitric oxide synthases. (7) Peroxynitrite ONOO is produced by the reaction of the free radical superoxide O2●− with the free radical nitric oxide NO. (8) Hypochlorous acid (HOCl) is produced through myeloperoxidase (MPO)-catalysed peroxidation of chloride anions using H2O2. (9) Nitrogen dioxide (NO2) reacts with molecular oxygen to form, under the action of UV radiation and heat, nitric oxide (NO) and ozone (O3). Lipid peroxidation (10)–(12) is a chain of reactions of oxidative degradation of lipids: (10) unsaturated lipid (LH) reacts with prooxidants (hydroxyl radical, HO), leading to the formation of lipid radical (L) and water; (11) lipid radical then reacts with oxygen to form a lipid peroxy radical (LOO), which abstracts one hydrogen from another lipid molecule, generating a new lipid radical (L) and lipid hydroperoxide (LOOH) (12). (C) The redox balance is an equilibrium between ROS production, due to pro-oxidant conditions, and antioxidant defences. A disruption of this redox balance in favour of oxidative conditions will promote an oxidative stress. CAT, catalase; CoQ, coenzyme Q; Cyt C, cytochrome C; GPx, glutathione peroxidase; GR, glutathione reductase; GSH, reduced glutathione; GSSG, oxidized glutathione; SOD, superoxide dismutases; Trx, thioredoxin.
Antioxidants 11 01426 g001
Figure 2. NRF2-KEAP1 signalling pathway in basal conditions. Under homeostatic conditions, NRF2 is sequestered by cytoplasmic KEAP1/CUL3-RBX1 E3 ubiquitin ligase complex and targeted to proteasomal degradation. CUL3, Cullin-3; KEAP1, Kelch-like ECH-associated protein 1; NRF2, nuclear factor erythroid 2-related factor 2; RBX1, RING Box Protein 1; Ub, Ubiquitin.
Figure 2. NRF2-KEAP1 signalling pathway in basal conditions. Under homeostatic conditions, NRF2 is sequestered by cytoplasmic KEAP1/CUL3-RBX1 E3 ubiquitin ligase complex and targeted to proteasomal degradation. CUL3, Cullin-3; KEAP1, Kelch-like ECH-associated protein 1; NRF2, nuclear factor erythroid 2-related factor 2; RBX1, RING Box Protein 1; Ub, Ubiquitin.
Antioxidants 11 01426 g002
Figure 3. Activation and regulation of the NRF2 signalling pathway. Under stress conditions, NRF2 is released from KEAP1 and translocates to the nucleus, where it interacts with cofactors and binds specific response elements (ARE) to regulate the transcription of its many target genes. ARE, Antioxidant Response Element; ATG5, Autophagy-related 5; BACH1/2, BTB and CNC homology 1/2; BCL2, B-cell lymphoma 2; BCRP, Breast cancer resistance protein; BRCA1, Breast cancer type 1; CAT, Catalase; CYP, Cytochrome p450; DPP3, Dipeptidyl peptidase 3; FTH1, Ferritin Heavy Chain 1; FTL, Ferritin light chain; FPN1, Ferroportin1; G6PD, Glucose-6-phosphate dehydrogenase; GCLC, Glutamate-cysteine ligase catalytic subunit; GCLM, Glutamate-cysteine ligase regulatory subunit; GPX2/8, Glutathione peroxidase 2/8; GSR, glutathione reductase; GST, glutathione S-transferase; HMOX1, Heme Oxygenase 1; IDH1, isocitrate dehydrogenase 1; MAO-A, monoamine oxidase A; ME1, malic enzyme 1; MRP1, Multidrug resistance protein 1; NQO1, NAD(P)H:quinone oxidoreductase 1; NF-κB, Nuclear factor kappa-light-chain-enhancer of activated B cells; p21, cyclin-dependent kinase inhibitor 1; p62, sequestosome 1; PALB2, Partner and localizer of BRCA2; PGAM5, Phosphoglycerate mutase 5; PGD, 6-phosphogluconate dehydrogenase; ROS, Reactive oxygen species; sMAF, small musculoaponeurotic fibrosarcoma; SOD, Superoxide dismutase; THXR, Thioredoxin reductase; UGT, UDP-glucuronosyltransferase; WTX, Wilms tumour gene on the X chromosome.
Figure 3. Activation and regulation of the NRF2 signalling pathway. Under stress conditions, NRF2 is released from KEAP1 and translocates to the nucleus, where it interacts with cofactors and binds specific response elements (ARE) to regulate the transcription of its many target genes. ARE, Antioxidant Response Element; ATG5, Autophagy-related 5; BACH1/2, BTB and CNC homology 1/2; BCL2, B-cell lymphoma 2; BCRP, Breast cancer resistance protein; BRCA1, Breast cancer type 1; CAT, Catalase; CYP, Cytochrome p450; DPP3, Dipeptidyl peptidase 3; FTH1, Ferritin Heavy Chain 1; FTL, Ferritin light chain; FPN1, Ferroportin1; G6PD, Glucose-6-phosphate dehydrogenase; GCLC, Glutamate-cysteine ligase catalytic subunit; GCLM, Glutamate-cysteine ligase regulatory subunit; GPX2/8, Glutathione peroxidase 2/8; GSR, glutathione reductase; GST, glutathione S-transferase; HMOX1, Heme Oxygenase 1; IDH1, isocitrate dehydrogenase 1; MAO-A, monoamine oxidase A; ME1, malic enzyme 1; MRP1, Multidrug resistance protein 1; NQO1, NAD(P)H:quinone oxidoreductase 1; NF-κB, Nuclear factor kappa-light-chain-enhancer of activated B cells; p21, cyclin-dependent kinase inhibitor 1; p62, sequestosome 1; PALB2, Partner and localizer of BRCA2; PGAM5, Phosphoglycerate mutase 5; PGD, 6-phosphogluconate dehydrogenase; ROS, Reactive oxygen species; sMAF, small musculoaponeurotic fibrosarcoma; SOD, Superoxide dismutase; THXR, Thioredoxin reductase; UGT, UDP-glucuronosyltransferase; WTX, Wilms tumour gene on the X chromosome.
Antioxidants 11 01426 g003
Figure 4. NRF2-signalling-based therapeutic strategies in brain dysfunctions. Targeting NRF2 pathway in brain diseases appears as a promising therapeutic strategy. Different options exist depending on the context (cancer (A), ageing (B), and neurodegenerative diseases (C)), NRF2 levels, and the cells to target (neuronal or glial cells). NRF2 and NRF2-negative regulators (KEAP1, BACH1, BACH2) can be targeted in these different conditions. The emergence of senotherapeutic drugs (senolytics, senomorphics) coupled to epigenetic modifiers (inhibitors/activators of histone acetyltransferases, histone deacetylases, histone demethylases, and histone methyltransferases) could provide a great number of translational opportunities and thus contribute to improving human health.
Figure 4. NRF2-signalling-based therapeutic strategies in brain dysfunctions. Targeting NRF2 pathway in brain diseases appears as a promising therapeutic strategy. Different options exist depending on the context (cancer (A), ageing (B), and neurodegenerative diseases (C)), NRF2 levels, and the cells to target (neuronal or glial cells). NRF2 and NRF2-negative regulators (KEAP1, BACH1, BACH2) can be targeted in these different conditions. The emergence of senotherapeutic drugs (senolytics, senomorphics) coupled to epigenetic modifiers (inhibitors/activators of histone acetyltransferases, histone deacetylases, histone demethylases, and histone methyltransferases) could provide a great number of translational opportunities and thus contribute to improving human health.
Antioxidants 11 01426 g004aAntioxidants 11 01426 g004b
Table 1. Methods for NRF2 activity monitoring.
Table 1. Methods for NRF2 activity monitoring.
ActivitiesDescriptionMethodsPrincipleObservationsRefs
Interaction KEAP1-NRF2The complex NRF2-KEAP1 is sequestered in the cytosol before being rapidly ubiquitinated and degradated by the proteasome.Fluorescence Lifetime Imaging–Förster Resonance Energy Transfer (FLIM-FRET) approach in cells expressing fluorescently tagged NRF2 and KEAP1evaluation of conformation changes in KEAP1-NRF2 complex in cells/establishment of NRF2 activation kineticsConformational changes (spatial distribution) can result in functional inactivation of the complex and consequential induction of an NRF2-mediated stress response.[84,85,86]
NRF2 translocationFree NRF2 will enter in the nucleus and activate the transcription of NRF2 target genes.Western blotsevaluation of the NRF2 localisation (cytoplasmic versus nuclear levels)increase in NRF2 level in the nucleus[87,88,89]
immunofluorescence (cells, tissues)NRF2 staining in the nucleusincrease in NRF2 level in the nucleus[87,90]
transcriptomic analysesanalyses of NRF2 target genes expressionincrease in gene expression[53,87,91]
NRF2 binding to AREsFree NRF2 enters in the nucleus and binds to specific sequences (Antioxidant Response Elements, ARE). This binding will activate the transcription of genes that have ARE sequences on their promoter.ARE luciferase reporter kitCells are transfected by an ARE luciferase reporter vector constitutively expressing Renilla luciferase vector.Amount of light generated after treatments is positively correlated with the ARE luciferase reporter activity as a consequence of NRF2 binding to ARE sequence.[92]
Electrophoretic-Mobility Shift Assay (EMSA)Nuclear extracts are incubated with an NRF2 probe (oligonucleotide containing the ARE consensus sequence). Protein/DNA complexes are separated on a nondenaturing polyacrylamide gel and transferred to a membrane. Detection is performed using Streptavidin-HRP Conjugate and a chemiluminescent substrate.The shifted bands corresponding to the protein/DNA complexes (interaction of NRF2 with its ARE-probe) are visualized.[93]
DNA-binding ELISA for activated NRF2 transcription factorNRF2 transcription factor (from nuclear extracts) binds to DNA sequence (containing the NRF2 consensus binding site) immobilized in the well. Incubation with primary and secondary antibodies specifically quantifies the amount of activated transcription factor.This method is 100-fold more sensitive than EMSAs. Colorimetric readout enables easy, quantitative analysis by spectrophotometry at 450 nm.[94]
NRF2/ARE luciferase reporter stable cell linesAfter treatments, the NRF2 transcription factor will enter in the nucleus and bind to its response element. As a consequence, luciferase is expressed, and light will be generated in an enzymatic assay (addition of luciferin).Amount of light generated is positively correlated with the level of NRF2 activation.[95]
high-content imaging of cell lines expressing fluorescent-tagged NRF2 (fluorescent protein reporter cell lines, expressing GFP tagged Nfe2l2)High-throughput live confocal imaging is used to measure the temporal dynamics of the NRF2 pathway after treatments.Fluorescent protein signal intensity is correlated to NRF2 response.[96]
GSH productionNRF2 regulates GSH biosynthesizing enzymes (GCLM, GCLC) and plays a key role in the regulation of cellular GSH homeostasisGSH quantification (GSH assay kit)GSH is oxidized by the sulfhydryl reagent 5,5’-dithio-bis(2-nitrobenzoic acid). The formed derivative is measurable at 412 nm. The glutathione disulphide (GSSG) formed can be recycled to GSH by glutathione reductase in the presence of NADPH.Increase in the GSH levels may confirm NRF2 recruitment. Glutathione quantification can be performed in any biological fluids, tissues, and cell extracts.[97]
Nfe2l2manipulationactivation/inhibition of the NRF2 signalling pathway and inactive-luciferase-tagged Nfe2l2, can be set up in transgenic mouse modelstransgenic mouse modelsTransgenic NRF2−/− mice (NRF2 pathway is downregulated) or KEAP1−/− mice (NRF2 pathway is upregulated) can be useful in order to verify the implication of the NRF2 pathway during pathologies or ageing.Transgenic animal models are commercially available.[98,99]
OKD48 transgenic miceThe OKD48 construct has a 3xARE promoter, human Nfe2l2, and Flag-tagged luciferase. Upon stress, OKD48 is transcriptionally induced by the 3xARE element, and luciferase activity (luminescence) is observed only in cells experiencing oxidative stress.OKD48-luc model is a highly specific and sensitive system for screening NRF2 activity[100,101]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Heurtaux, T.; Bouvier, D.S.; Benani, A.; Helgueta Romero, S.; Frauenknecht, K.B.M.; Mittelbronn, M.; Sinkkonen, L. Normal and Pathological NRF2 Signalling in the Central Nervous System. Antioxidants 2022, 11, 1426. https://doi.org/10.3390/antiox11081426

AMA Style

Heurtaux T, Bouvier DS, Benani A, Helgueta Romero S, Frauenknecht KBM, Mittelbronn M, Sinkkonen L. Normal and Pathological NRF2 Signalling in the Central Nervous System. Antioxidants. 2022; 11(8):1426. https://doi.org/10.3390/antiox11081426

Chicago/Turabian Style

Heurtaux, Tony, David S. Bouvier, Alexandre Benani, Sergio Helgueta Romero, Katrin B. M. Frauenknecht, Michel Mittelbronn, and Lasse Sinkkonen. 2022. "Normal and Pathological NRF2 Signalling in the Central Nervous System" Antioxidants 11, no. 8: 1426. https://doi.org/10.3390/antiox11081426

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