Next Article in Journal
Phoenix dactilyfera L. Pits Extract Restored Bone Homeostasis in Glucocorticoid-Induced Osteoporotic Animal Model through the Antioxidant Effect and Wnt5a Non-Canonical Signaling
Next Article in Special Issue
The Involvement of Sirtuin 1 Dysfunction in High-Fat Diet-Induced Vascular Dysfunction in Mice
Previous Article in Journal
Vitreous Humor Proteome: Targeting Oxidative Stress, Inflammation, and Neurodegeneration in Vitreoretinal Diseases
Previous Article in Special Issue
Pequi Fruit Extract Increases Antioxidant Enzymes and Reduces Oxidants in Human Coronary Artery Endothelial Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Genetic Variations on Redox Control in Cardiometabolic Diseases: The Role of Nrf2

by
Cecilia Zazueta
1,
Alexis Paulina Jimenez-Uribe
2,
José Pedraza-Chaverri
2 and
Mabel Buelna-Chontal
1,*
1
Departmento de Biomedicina Cardiovascular, Instituto Nacional de Cardiología, I.Ch., Mexico City 14080, Mexico
2
Departamento de Biología, Facultad de Química, Universidad Nacional Autónoma de México, Mexico City 04510, Mexico
*
Author to whom correspondence should be addressed.
Submission received: 10 February 2022 / Revised: 1 March 2022 / Accepted: 4 March 2022 / Published: 6 March 2022
(This article belongs to the Special Issue Oxidative Stress and Inflammation in Cardiovascular Diseases II)

Abstract

:
The transcription factor Nrf2 is a master regulator of multiple cytoprotective genes that maintain redox homeostasis and exert anti-inflammatory functions. The Nrf2-Keap1 signaling pathway is a paramount target of many cardioprotective strategies, because redox homeostasis is essential in cardiovascular health. Nrf2 gene variations, including single nucleotide polymorphisms (SNPs), are correlated with cardiometabolic diseases and drug responses. SNPs of Nrf2, KEAP1, and other related genes can impair the transcriptional activation or the activity of the resulting protein, exerting differential susceptibility to cardiometabolic disease progression and prevalence. Further understanding of the implications of Nrf2 polymorphisms on basic cellular processes involved in cardiometabolic diseases progression and prevalence will be helpful to establish more accurate protective strategies. This review provides insight into the association between the polymorphisms of Nrf2-related genes with cardiometabolic diseases. We also briefly describe that SNPs of Nrf2-related genes are potential modifiers of the pharmacokinetics that contribute to the inter-individual variability.

Graphical Abstract

1. Introduction

Worldwide cardiometabolic diseases-related deaths have been increasing over the last decades, representing a big concern [1]. As cardiometabolic diseases lead to unbalance of redox homeostasis and enhanced oxidative stress (for a recent comprehensive review, see [2]), different antioxidant therapies have been evaluated to prevent and treat such diseases. Although some of these have been designed to induce the nuclear factor erythroid 2-related factor 2 (Nrf2)-driven antioxidant response [3], their efficacy in the clinic has not been clearly established, giving rise to the possibility of the participation of other regulatory elements in the Nrf2-signaling cascade. In this regard, multiple single nucleotide polymorphisms (SNPs) have been found in the Nrf2 gene [4]. Overall, SNPs of the Nrf2 gene induce changes in Nrf2 activation, which can lead to the impairment of the endogenous antioxidant system, contributing to chronic inflammation and other detrimental effects [5]. Indeed, several polymorphisms of Nrf2 have been associated with cardiometabolic diseases progression [6]. This review highlights the association of SNPs of Nrf2 and related genes with cardiometabolic diseases that significantly impact disease progression, as well as a short look at the possible implications on cardiovascular drug therapy.

2. The Role of Nrf2 to Maintain Redox Homeostasis in Cardiometabolic Diseases

A common feature of cardiometabolic diseases is the imbalance between pro- and anti-oxidative factors in the cell. Such condition is associated with high levels of reactive oxygen species (ROS) and reactive nitrogen species (RNS) that react with lipids, proteins, DNA, or activate redox signaling pathways, leading to cellular injury and death. Disruption of the redox homeostasis is associated with poor cardiac contractility related to reduced sarco-endoplasmic reticulum Ca2+-ATPase (SERCA2A) activity [7], but also with mitochondrial damage. Mitochondria are the primary ROS producers and, at the same time, are the first exposed in front line due to their deleterious action, resulting in further mitochondrial dysfunction and a vicious cycle of ROS generation. Impairment of the mitochondrial function and increased rates of fatty acid β-oxidation promote the accumulation of incomplete β-oxidation products, which, in combination with oxidative stress, contribute to insulin resistance in cardiometabolic diseases [8].
In cardiometabolic diseases, among the antioxidant molecules regulated by Nrf2 are catalase (CAT), superoxide dismutase (SOD), glutathione S-transferases (GST), glutathione peroxidases (GPx), heme-oxygenase 1 (HMOX-1), thioredoxin (TXN), thioredoxin reductases, peroxiredoxins (Prdx), and NAD(P)H quinone oxidoreductase-1 (NQO1) [9,10,11,12] (Figure 1A). Besides the antioxidant response regulation, Nrf2 also modulates a myriad of genes related to other cellular processes. For instance, it is known that Nrf2 regulates the expression of Notch1 and downstream genes, which collectively coordinate signals that affect differentiation, proliferation, and apoptotic events [13]. Moreover, it was demonstrated by chromatin immunoprecipitation that Nrf2 binds directly to an antioxidant response elements (ARE) region in the aryl hydrocarbon receptor (AHR) promoter [14]. AHR increased in pig myocardium after ischemia-reperfusion damage and diminished upon administration of natural flavone baicalin [15] and sulforaphane [16] in association with lesser cardiac injury and inflammation.

Nrf2 Structure and Regulatory Mechanisms

Nrf2 belongs to the cap‘n’ collar (CNC) subfamily of the basic leucine zipper (bZip) transcription factors [17]. Nrf2 contains seven domains (Figure 1B) named Nrf2-ECH homology (Neh) 1 to 7 [18,19]. Among Neh domains, Neh1 and Neh2 have the most recognizable roles in Nrf2 function and regulation. Neh2 domain is located at the amino-terminal and has ETGE and DLG motifs that interact with Kelch-like ECH-associated protein 1 (Keap1), the well-known negative regulator of Nrf2 [20]. Neh4 and Neh5 are two domains that allow Nrf2 interaction with other proteins such as co-activators [21,22,23,24], whereas Neh5 possesses a nuclear export signal (NES) sensitive to oxidation [25]. Neh7 and Neh6 domains act as negative regulators, and Neh7 interacts with the retinoic X receptor alpha (RXRα) to repress Nrf2 transcriptional activity. In fact, RXRα competes with Nrf2 for binding to the ARE sequence [19]. On the other hand, Neh6 interacts with beta-transducin repeat-containing protein (β-TrCP) to promote Nrf2 ubiquitination and consequent degradation [26,27]. Neh1 domain located near the C-terminal region harbors a basic leucine zipper motif, which binds to ARE sequences within target genes. Neh1 also interacts with small musculoaponeurotic fibrosarcoma (sMAF) proteins favoring the transcription of antioxidant genes. In addition, Neh1 acetylation enhances the transcriptional activity of Nrf2 [28,29]. The Neh3 domain in the C-terminal region interacts with chromo-ATPase/helicase DNA binding protein 6 (CDH6), essential for Nrf2 transcriptional activity [30].
Under homeostatic conditions, Nrf2 is maintained at low levels due to constant proteasomal degradation. The degradation process occurs through the interaction of the ETGE and DLG motifs within the Nrf2 Neh2 domain with a Keap1 homodimer [20]. Keap1 acts as an adaptor for an E3 ubiquitin ligase complex that contains cullin 3 (CUL3) and ring box-1 (Rbx) [31], which catalyze the ubiquitination of lysine residues located between ETGE and DLG motifs [32,33] to induce Nrf2 degradation through the ubiquitin proteasomal system (UPS). Interestingly, Nrf2 also promotes CUL3 and Rbx expression as negative feedback [34]. Another mechanism that induces Nrf2 degradation via UPS is Neh6 phosphorylation by glycogen synthase kinase 3 (GSK-3); this promotes the binding of β-TrCP, which acts as a substrate for the Skp1–Cullin–F-box protein complex (SFC), another E3 ubiquitin ligase complex (Figure 2A) [35]. Of note, the levels of GSK-3 increase in experimental models of obesity, diabetes, and associated cardiovascular diseases [36,37,38].
During oxidative stress, ROS can react with Keap1 cysteine residues modifying the Keap1 homodimer structure, allowing Nrf2 release. Hence, Keap1 also acts as a sensor of redox homeostasis [39,40,41,42,43]. Once Nrf2 is released from the Keap1/Cul3/Rbx complex, nuclear localization signals (NLS) are exposed for importins recognition to elicit Nrf2 nuclear translocation [44]. After heterodimerization with sMAF, Nrf2 induces the expression of antioxidant response-related genes [45]. Another potential mechanism for Nrf2 activation involves its phosphorylation by several kinases. Protein kinase C (PKC) phosphorylates Nrf2 at serine 40 of the Neh2 domain, inducing its nuclear translocation [46]. Casein kinase 2 (CK2) can phosphorylate Neh4 and Neh5 transactivation domains, enhancing Nrf2 transcriptional activity [47]; in addition, CK2 phosphorylates AMP-activated protein kinase (AMPK) [48], which, in turn, phosphorylate serine residue 558 to increase Nrf2 nuclear accumulation [49], and serine residues 374, 408, 433 to increase its transcriptional activity [50]. The regulatory mechanisms of Nrf2 are depicted in Figure 2.

3. Nrf2 Genetic Variations (Single Nucleotide Polymorphisms)

Human Nrf2 gene (NFE2L2 gene ID: 4780) codes for Nrf2 mRNA 2859 base pairs long, which is translated into a protein containing 605 amino acid residues. Genome-wide association studies (GWAS) have been used to find sequence mutations of the Nrf2 locus including exonic SNPs [6]. Certainly, SNPs of the promoter region of the Nrf2 and ARE-related genes regulate cardiometabolic disease progression, as we will detail later. To date, multiple SNPs of the Nrf2 human gene have been found [4,6], including −617 C/A (rs6721961) and −651 G/A (rs6706649) in the promoter region, which attenuate Nrf2 transcriptional activity, diminishing ARE binding [51,52]. Likewise, the polymorphisms −686 A/G, −684 G/A, and −650 C/A and one triplet repeat polymorphism, −20 to −6 (CCG) 4 or 5, alter the basal expression of Nrf2, resulting in the increased severity of implicated diseases [5]. In the setting of chronic kidney disease, Nrf2 (−617 C/A) rs6721961, MnSOD (SOD2) rs4880, and GPx1 rs1050450 polymorphisms have been described as predictors of overall survival. Both rs6721961 and rs4880 impair antioxidant defense affecting survival, whereas rs1050450 contributes to longer cardiovascular survival in patients undergoing hemodialysis [53]. Furthermore, the polymorphism −653 A/G (rs35652124) in the Nrf2 gene was related to a higher inflammatory profile in liver samples from alcoholic liver disease patients [54]. Additionally, the Nrf2 rs35652124 and KEAP1 rs1048290 variants have been associated with chronic inflammatory diseases such as chronic obstructive pulmonary disease [55]. In this context, genetic variations in KEAP1 rs11085735 and Nrf2 rs2364723 can affect lung function [56]; in particular, the KEAP1 rs11085735 variant has been proposed as an independent strong predictor for incidence of cardiovascular events [57]. Furthermore, the functional polymorphism in promoter of Nrf2 gene (−617 C/A) is associated with a higher risk to develop oxidative stress-related and inflammatory illnesses, including acute lung injury [58]. An interesting 18-year follow-up study in 1390 subjects from the Vlagtwedde–Vlaardingen cohort, a general Caucasian population-based cohort, demonstrated that Nrf2 is associated with human survival. Carriers of the minor allele rs13001694 had less risk of all-cause mortality, whereas those carriers of rs2364723 showed a reduced risk for cardiovascular mortality and lower triglyceride levels [59].
Polymorphisms in the Nrf2 gene and related genes that decrease Nrf2 activation might be a risk for several diseases. Additionally, human aging is a detrimental condition, where Nrf2 activation declines due to genetic variations. Elder patients carrying the rs35652124 variant in the Nrf2 gene show frailty, adverse drug reactions, and multimorbidity [60]. The decreased Nrf2 activation in aging concurs with unbalance of redox homeostasis, supporting the theory that oxidative stress and the proinflammatory state contribute to aging and to an increase in cardiovascular diseases. Hence, either genetic or non-genetic factors leading to the impairment of Nrf2 transcriptional activity, including Nrf2 de novo synthesis, not only cause a faulty endogenous antioxidant system but also contribute to the development of several chronic inflammatory diseases [54].

4. Contribution of Nrf2/Keap1 and Target Genes Polymorphisms in Cardiometabolic Diseases

4.1. Obesity and Diabetes

Obesity is defined as an imbalance between energy intake and energy expenditure resulting in adipose tissue hyperplasia; however, this pathology involves a complex network of metabolic alterations at cellular, tissular, and systemic levels that lead to low-grade inflammation and oxidative stress. In addition, this pathology is the hub for developing other non-communicable pathologies such as type 2 diabetes mellitus (T2DM), hypertension, steatohepatitis, and metabolic syndrome. Recently, obesity has been described as a low-grade state of chronic inflammation that also drives to an oxidative stress status in different organs besides adipose tissue; it is well known that both inflammation and oxidative stress can promote insulin resistance and T2DM development [61]. Redox imbalance occurs in obese [62,63,64,65] and in diabetic patients [66]; particularly in T2DM, oxidative stress is potentiated due to the presence of other inducers such as advanced glycation products derived from chronic hyperglycemia [67]. Hence, a great variety of natural Nrf2 activators has beneficial effects in experimental models of obesity and T2DM, showing a reduction in the inflammatory response, decrease in body weight gain, and lesser clinical complications [68,69,70,71,72,73,74,75,76,77]. The abovementioned supports the crucial role of Nrf2 in preventing the metabolic derangements associated with obesity and T2DM. Lower Nrf2 activity in experimental models of insulin resistance and diabetes has been found in different organs [78,79]. In addition, genetic variation contributes to Nrf2 dysfunction in obesity and in T2DM development, as we will detail further.
Nrf2 polymorphism rs2364723 has been correlated with changes in heart rate variability through an imbalance of autonomous nervous system in overweighted patients (SAPALDIA cohort), impairing cardiovascular function [80]. Moreover, polymorphisms found in Nrf2-related genes including SOD1 (rs2234694), SOD3 (rs2536512), GSTM1 (rs1056806), SOD2 (rs4880), and GPx1 (rs1800668) were prevalent in obese patients, and the last four were associated with increased risk of obesity [81]. Another study found a positive association between elderly obese subjects and the SOD2 rs4880 polymorphism, independent of other factors [82]. It was suggested that this polymorphism attenuates SOD2 activity and increases oxidative stress [83]. In addition, rs4880 was associated with higher oxidized-LDL (ox-LDL) levels and T2DM prevalence [84]. In this regard, a study in Japanese–American subjects indicates that SOD2 rs4880 polymorphism might be associated with the progression of T2DM supported by observed oxidative stress, which leads to glucose intolerance [85]. On the other hand, attenuated SOD2 activity in those patients carrying the rs4880 polymorphism affects not only redox homeostasis but also increases the cardiovascular risk associated with T2DM and inflammatory response related to the ox-LDL, which contributes to the onset and progression of atherosclerosis [84]. Although the valine/alanine polymorphism of SOD2 (rs4880) was not associated with diabetes etiology, it was associated with diabetic nephropathy in Japanese patients with T2DM [86]. Moreover, the rs4880 variant increased the risk for diabetic nephropathy in type 1 diabetes patients [87,88] and is a predictor of cardiovascular disease [88]. The HMOX1 gene is associated negatively with diabetic complications, since the rs2364723 CG or CC allele increased in serum of T2DM patients compared to controls [89]. Additionally, the GPx rs1050450 has been associated with morbid obesity development in Mexican population, particularly in females [90]. This polymorphism is also related with diabetic neuropathy in a Polish and English population [91,92], as well as with the development of carotid plaques in a Chinese population with diabetes [93].
Among the polymorphisms on the Nrf2 gene, seven are pointed out as relevant in T2DM and related complications. The polymorphism rs6721961 in the promoter region of the Nrf2 gene was associated with T2DM development in Mexican patients [94], as well as diabetic complications such as diabetic nephropathy in Chinese patients [95]. Indeed, rs6721961 was strongly associated with a higher risk for the progression of T2DM in a Chinese population. The authors proposed that this polymorphism may impair β-cell function and increase insulin resistance, contributing to the onset and progression of T2DM [96]. Furthermore, the rs35652124 polymorphism, also located at the Nrf2 promoter region, seems to be a harmful genetic variant that predisposes subjects to insulin resistance and impaired angiogenesis, associated with diabetic foot ulcer development, in an Indian population [97]. The Nrf2 rs182428269 variant has been associated with T2DM risk and diabetic-associated complications in the Indian population [98]. On the other hand, the Nrf2 polymorphisms rs2364723, rs10497511, rs1962142, and rs6726395 have not been associated with the risk of developing T2DM but were significantly associated with diabetic complications in Chinese patients [89]. A recent study suggests that Nrf2 expression-associated variants are likely to be a useful indicator of T2DM development in the human population [99].
The above reported suggest that Nrf2 and -related genes polymorphisms are associated with obesity development, diabetes prevalence and progression, as well as diabetes-related complications, highlighting the role of redox homeostasis in diabetes pathophysiology.

4.2. Coronary Artery Disease (CAD)

CAD is a multifactorial chronic disease associated with several risk factors such as hypertension, diabetes, obesity, dyslipidemia, smoking, and genetics. The major cause of CAD is atherosclerosis, which results from the complex interplay between dyslipidemia, endothelial dysfunction, and oxidative stress. In the last decade, it has become evident the relationship between reduced Nrf2 activity and the development of cardiovascular disorders in patients with obesity, diabetes mellitus, and atherosclerosis [100,101], confirming a plethora of studies using pathological models in isolated cells as well as in animals, where the low activity of the Nrf2-Keap1-ARE axis correlates with augmented oxidative stress and detrimental cellular effects. The regulatory mechanisms of these pathways have been reviewed in depth [102,103]. However, the information about the genetic variability on this pathway among individuals, groups, or populations is still scarce.
The survival analysis of the Vlagtwedde–Vlaardingen cohort, a general population-based cohort of Caucasian individuals of Dutch descent recruited from both a rural area and an urban area in the Netherlands, supports the hypothesis that Nrf2 may be a gene that contributes to individual differences in human lifespan. The A/G substitution in the rs13001694 variant may occur only under oxidative stress conditions (smoking habits) and was associated with a lower risk of all-cause mortality. Those minor allele carriers of rs2364723 are associated with a reduced risk of cardiovascular mortality within groups with different risk factors, i.e., gender and smoking. The rs2364723 prevalence was associated with lower triglyceride levels in male subjects, supporting that this Nrf2 variant contributes to the lower cardiovascular mortality observed [59]. Conversely, the rs35652124 polymorphism was associated with cardiovascular mortality in Japanese patients subjected to hemodialysis. A controversial issue in this study was that the G allele carriers of the rs35652124 variant, which is, in turn, associated with a diminished Nrf2 function, showed low cardiovascular mortality [104]. A possible explanation is that such an effect occurs through Nrf2 regulation on plasma lipoproteins and cholesterol, which overwhelms its antioxidant function. Studies in Nrf2 knock-out mice exhibited greater oxidative stress and concurred with minor aortic atherosclerosis, sustaining such a hypothesis [105]. Some of the mechanisms involved include decreased CD36 expression, and thereby decreased LDL uptake and reduced cholesterol-dependent inflammasome activation [106]. Recently, using two different murine models of atherosclerosis, LDL receptor-deficient mice (LDLR/−) and LDLR/− mice expressing only apoB-100 (LDLR/−ApoB100/100), demonstrated that Nrf2 deficiency delays early atherogenesis but, in later stages of the disease, promotes plaque instability [107]. Additionally, in a study that included 2374 Thai subjects with a high risk of CAD and without CAD, the association between Nrf2 rs6721961 TT, the severity of coronary atherosclerosis, and CAD was demonstrated in the entire population, whereas the association with the NQO1 rs1800566 CC genotype was established only in female [108].
Other studies have focused on studying variations in genes involved in oxidative stress at early atherosclerosis. For instance, p66Shc is an Nrf2-regulated gene [109] whose deletion correlated with lower oxidative stress, as well as reduced early atherogenic lesions in knockout p66Shc/− mice under a high-fat diet [110]. The p66Shc genetic variations, −354T>C and 92C>T, are scarce in subjects at the early onset of coronary artery disease, although these gene variants are not associated with the risk for CAD [111]. On the other hand, due to the risk of coronary events in chronic kidney disease patients being comparable with those with diabetes, the association of the genetic variability of the Nrf2-Keap1 axis in cardiovascular diseases has also been reported in those patients. In chronic kidney disease patients, a cohort carrying the A allele in the rs11085735 polymorphism located in the proximity of exon 3 of KEAP1 was defined as an independent predictor of incident cardiovascular events [57].
Studies of SNPs of genes that contain ARE sequence in their promoters are scarce. Data from integrated computational analysis, chromatin immunoprecipitation sequencing (ChIP-Seq), and GWAS have been fundamental to identify some functional SNPs of ARE-containing genes [112]. Wang et al. [113] reported that the polymorphic ARE rs242561 C/T in MAPT gene (encoding microtubule-associated protein Tau) showed a strong affinity for Nrf2/sMAF heterodimers [113], likely inducing diverse responses in diseases in which conditions of oxidative stress prevail. Another example is the SNP rs113067944 within the ARE sequence of the ferritin light polypeptide (FTL) gene promoter. Changes in allele A to C in the core ARE region TG[A/C]CTCAGC decreased the Nrf2 binding affinity, resulting in a lower FTL transcription. In the study conveyed by Kuosmanen and coworkers [114], 14 different SNPs of the ARE sequences with putative clinical relevance have been described. Cardiovascular diseases were not considered in this analysis; however, it is well known that genes implicated in the metabolism of intracellular iron and ferroptosis, including ferritin light chain 1, are associated with cardiomyopathy and related diseases, such as myocardial infarction [115]. Conversely, the AA genotype of HMOX1 was associated with the diminished incidence of ischemic heart disease in a Japanese study that included 1972 control subjects and 597 patients with myocardial infarction [116]. On the other hand, the role of the rs1800566 T carrier genotype (CT + TT) in the NQO1 gene is controversial. Whereas Han et al. [117] have established an association with a higher risk to develop carotid artery plaques, other groups have reported that NQO1 rs1800566 (CT + TT) has a protective effect on CAD risk [118]. Of note, there is a clear correlation between allele dosage and enzymatic activity in NQO1 rs1800566, i.e., the homozygotes (TT) have the lowest, the heterozygotes (CT) intermediate, and the wild-type homozygotes (CC) the highest NQO1 enzyme activity [119]. Interestingly, an association was found between the Val16Ala-SOD2 polymorphism of the SOD2 gene and oxidative stress injury in hypercholesterolemic patients [120]. We need a better comprehension of how SNPs of the Nrf2 and target genes contribute to the occurrence and progression of CAD to perform efficient redox-targeted therapy to treat cardiovascular diseases.

4.3. Hypertension

Nrf2-Keap1-ARE axis is fundamental in the antioxidant defense response in cardiovascular diseases, including atherosclerosis and hypertension [121]. Polymorphisms of genes coding for antioxidant regulators and enzymes might contribute to the modulation of their function in both conditions. As we mentioned before, polymorphisms of the promoter region of the Nrf2 gene might cause lower transcription, activity, and related gene expression. One of these SNPs, the rs6721961 (G>T) variation, was associated with elevated systolic and diastolic blood pressure levels and with cardiovascular mortality in Japanese patients undergoing hemodialysis, especially in the female cohort [104]. This finding was later confirmed by Kunnas et al. [122] in the TAMRISK study, in which hypertensive subjects were compared with normotensive ones in a Finnish cohort of 50-year-old subjects, followed up for ten years. They demonstrated that rs6706649 polymorphism located in the 5′ regulatory region of Nrf2 was not associated with cardiovascular disease. Endothelial dysfunction is a well-known factor implicated in hypertension. In this regard, individuals carrying the Nrf2 polymorphisms -617A and -653G showed impaired forearm vasodilator responses, affecting their vascular function [123].
HMOX1 gene has at least two polymorphic sites in its promoter at the 5′-flanking region, the rs3074372 (GT)n dinucleotide repeat, and rs2071746 T/A [124]. The A allele has been associated with HMOX1 higher transcription rates and higher blood pressure in women [116]. Moreover, the SOD3 c.172G>A variant and CAT c.-20C>T influenced blood pressure values and were associated with the risk for hypertension. Likewise, the polymorphisms of TXN gene c.-793T>C and GPx1 gene c.*891C>T also influenced blood pressure values [125]. Finally, the NQO1 rs1800566 CC genotype and HMOX1 rs2071746 showed marginal association with hypertension that increased when the combined effect was explored [108]. Therefore, this suggests that different genetic variants might interact or have synergic effects, resulting in increased susceptibility for CAD development risk factors. The SNPs of Nrf2 and related genes associated with cardiometabolic diseases are shown in Table 1.

5. Implications of Nrf2-Related Genes Polymorphisms in Cardiovascular Drug Therapy

Genetic polymorphisms cause interindividual variability that impacts the pharmacokinetics of anti-platelet, anti-coagulant, statins, and other drugs used for cardiovascular treatment, impairing their efficacy. Here, we briefly focus on some genetic polymorphisms of antioxidant-related genes.
Human UGT (UDP-glucuronosyltransferase)-2B7 regulated transcriptionally by Nrf2 is crucial for the metabolism of several drugs. Among other compounds, UGT drives phase II glucuronidation of the anticoagulant warfarin and its metabolite, hydroxywarfarin. Stable doses of warfarin in patients with mechanical cardiac valves were different among genotypes. In particular, the patients with the UGT1A1 (rs887829) variant required higher doses of warfarin [126]. Genetic polymorphisms of the UGT2B7 promoter region impair Nrf2-mediated transcription altering the drug efficacy [127]. On the other hand, the antiplatelet clopidogrel requires cytochrome P540 enzymes (CYPs) for its biotransformation into thiol active metabolites. Genetic variations of CYP2C19 (*2 and *3 alleles) exert inter-individual variability in the pharmacokinetics of clopidogrel dosing [128]. Conversely, in patients with acute coronary syndrome, the polymorphism Q192R of paraoxonase-1 enzyme (PON1), which codes for a high-density lipoprotein-associated antioxidant enzyme (containing ARE/Nrf2 binding sites in its promoter), did not hamper the biotransformation of clopidogrel, unaltering its potential efficacy to treat cardiovascular events [129]. A common cardiovascular treatment is the use of statins to lower circulating total cholesterol (TC) and low-density lipoprotein cholesterol (LDL-C). Rosuvastatin is a statin used in clinical trials for LDL-cholesterol lowering, which also decreases ROS production, inflammation, and induces an upregulation of antioxidant enzymes [83]. Therefore, rosuvastatin maintains both redox homeostasis and vascular function, preventing atherogenesis. Moreover, hypercholesterolemic patients carrying the VV genotype of the Val16Ala-SOD2 polymorphism showed low SOD2 activity and resistance to rosuvastatin therapy, with a concomitant increase in inflammatory and fibrinolytic biomarkers [83]. On the other hand, the CAT gene polymorphism CAT-262C>T in dyslipidemic patients was not associated with insulin resistance, blood lipids increase, or response to atorvastatin [130]. Recently, polymorphisms of SELENOP (selenoprotein P), rs3877899, and rs7579 were associated with differential responses to selenium, whose accumulation alters redox homeostasis and lipid profile in patients after Brazil nut consumption and statins treatment [131].
Based on this evidence, it is clear that inter-individual variation defined by genetic polymorphisms may determine drug efficacy and particular doses for cardiometabolic diseases’ treatment.

6. Coffee Consumption, Genetic Polymorphisms of Nrf2 and Cardiometabolic Diseases

We will briefly analyze the influence of coffee consumption on the risk of developing cardiometabolic diseases and its relation to polymorphisms of Nrf2-related genes. Coffee consumption has shown to be beneficial against cardiometabolic diseases. The acute effects of caffeine are increased metabolic rate/thermogenesis, insulin secretion, and reduced mean arterial pressure [132]. Indeed, moderate coffee intake in patients regulates lipid metabolism and prevents lipid oxidation and inflammation [133]. Recently, a meta-analysis discarded the idea that coffee or caffeine consumption is associated with the risk of atrial fibrillation [134]. Moderate coffee drinking is associated with a low risk of developing cardiometabolic diseases, such as T2DM [135], obesity [136,137,138], and heart failure [139]. Coffee-related beneficial effects may be mediated by coffee phytochemicals through an adaptive mechanism involving AHR, AMPK, sirtuins, and Nrf2-driven antioxidant genes transcription [135]. Conversely, high doses of caffeine can be detrimental to cardiovascular health by increasing blood pressure and heart rate [132]. Additionally, long-term high coffee consumption can increase serum triglycerides levels, total, and LDL-cholesterol, likely due to cafestol (a diterpene found in coffee), thus augmenting the risk for cardiometabolic diseases [140]. Certainly, the detrimental effects related to high coffee consumption involve a higher risk of CAD [141]. However, it has also been claimed that high coffee consumption is associated with lower risk for cardiovascular mortality [142]. These apparent contradictions might be associated with several factors, including genetic variability. On the other hand, coffee harvesters carrying the PON1 Q192R variant showed decreased PON1 activity, affecting the hydrolysis of pesticides and oxidized lipids, which caused increased intoxication and higher cardiovascular risk, associated with hypertension [143]. Accordingly, high PON1 enzymatic activity is associated with a lower cardiovascular risk [144].
In this regard, coffee intake induces Nrf2-dependent gene expression, although, depending on prevailing SNPs of the Nrf2 gene, high individual variability in the antioxidant response occurs [51]. A study showed that after four weeks of coffee intake, Nrf2 gene transcription increased in individuals carrying the SNPs rs6721961 (-617 C/A), rs35652124 (-653 A/G), and rs6706649 (651 G/A) [52]. Similarly, a two-month coffee consumption trial showed that in rs35652124 carriers, Nrf2, GST1A1, and UGT1A1 gene transcription increased compared to those carrying the rs6706649 genotype. Interestingly, before coffee consumption, the polymorphism rs35652124 was associated with lower Nrf2 transcription and further activation, whereas rs6706649 carriers showed higher Nrf2 gene transcription. These data suggest that patients with the Nrf2 rs35652124 variant, as well as with the B/B genotype in GST1A1 and [TA]6/6 or [TA]7/7 in the UGT1A1 gene could benefit more from coffee-mediated effects [145]. Notwithstanding, as we described earlier, although the Nrf2 rs6721961 and rs35652124 variants are detrimental for antioxidant defense in the setting of inflammatory diseases, rs6706649 has not been associated with cardiovascular diseases [53,55,121]. These data support the potential beneficial effects of coffee consumption in patients carrying these Nrf2 polymorphisms. Certainly, specific Nrf2 genetic variations can influence their ability to be activated by bioactive compounds [146]. Coffee consumption exerts a large influence on the Nrf2 pathway by individual genetic variation that might be related to the risk of cardiometabolic diseases.
Another intriguing issue is that gene variants that improve caffeine breakdown are expressed under high or prolonged coffee consumption. GWAS have identified an association between coffee consumption and the CYP (1A1/2) gene rs2472297 and rs2470893 variants, as well as AHR rs4410790 and rs6968865 variants [147,148,149]. It has been established that AHR cooperates with Nrf2 to regulate antioxidant gene expression [150,151,152]. Certainly, AHR senses xenobiotics, including polycyclic aryl hydrocarbons found in roasted coffee, also regulates the transcription of both CYP (1A1/2) [147]. CYPs enzymes are implicated in xenobiotics metabolism and are essential for caffeine metabolism [142]. Both rs2472297 and rs6968865 are associated with higher CYP activity, which enhances caffeine clearance and allows higher coffee consumption [147]. As we mentioned before, CYP and UGT enzymes are also essential for drug metabolism, including those used in cardiovascular therapy. Therefore, CYP and UGT polymorphisms might also directly influence pharmacokinetics, contributing to the inter-individual variability.

7. Association of Environmental Factors with Nrf2 and Related Genes Expression in Cardiometabolic Diseases

Environmental pollution includes fine and ultrafine particulate matter (PM). Some of the major components of PM include black carbon, elemental carbon, and organic carbon, which, in turn, might include polycyclic aromatic hydrocarbons, carbonyl compounds, n-alkanes, organic acids, and heterocyclic compounds [153,154]. Environmental pollution has been associated with a higher risk for cardiovascular diseases [155,156]. Since PM causes redox alterations and endothelial dysfunction [157], Nrf2 function seems to be highly relevant. A model of Nrf2 knock-out mice exposed to PM promoted vascular thickness, increased angiotensin II levels, and worsened cardiac function [158,159]. This phenomenon could also affect patients carrying Nrf2 polymorphisms that diminish its function. However, we only found one study in patients. Patients carrying the SNPs of GST rs1695, SOD2 rs4880, and Nrf2 rs1806649 showed higher susceptibility to PM and respiratory impairment due to air pollution [160]. Moreover, in elderly patients with cardiovascular diseases, the exposure of PM such as elemental carbon, black carbon, nitrogen oxides, and, in particular, organic carbon was associated with Nrf2 gene expression [161].
On the other hand, the gas generated from tobacco smoking contains fine PM and aldehydes, which are implicated in the loss of lung function [162]. Of note, tobacco smoking is also closely related to inflammation and endothelial dysfunction and is the leading factor of atherosclerosis development [163]. In this regard, the Nrf2 gene variant rs6721961 TT has been found in heavy cigarette smokers [164]. As mentioned earlier in this review, such Nrf2 gene variant is associated with T2DM prevalence, CAD, coronary atherosclerosis, elevated systolic and diastolic pressure levels, and cardiovascular mortality. Furthermore, a study in a Japanese population found that the Nrf2 rs6726395 polymorphism is associated with individual susceptibility to loss of lung function, owing to cigarette smoking [165]. Altogether, evidence support these environmental factors, such as air pollution and tobacco smoking, are related to Nrf2-related gene function and play a role in the risk of cardiovascular diseases.

8. Concluding Remarks

Numerous Nrf2 polymorphisms have been related to functional alterations associated with cardiometabolic diseases in diverse human populations. SNPs of Nrf2 and related genes contribute to susceptibility for obesity, inflammation, and diabetes progression, as well as coronary artery disease, hypertension, and cardiovascular mortality. Clearly, the endogenous antioxidant system deficiency is associated with several diseases; hence, the influence of the genetic variation in Nrf2 and ARE-containing genes that can impair the induction of the Nrf2-driven pathway accounts for susceptibility to suffer cardiometabolic diseases. Polymorphisms of Nrf2 and related genes have a clear association with the onset and risk of cardiometabolic diseases; therefore, they might be helpful to predict the incidence of cardiovascular events. Aside from their potential application as risk factors markers, Nrf2-target gene variants might be modifiers of pharmacokinetics contributing to the inter-individual variability and impairing the effectiveness of cardiovascular drug therapies. A better understanding of the implications regarding the polymorphisms of Nrf2 and related genes will allow us to design novel cardioprotective strategies to improve clinical outcomes due to complications in responses to therapies among individuals. Although we here describe the influence of Nrf2 gene variants on cardiovascular diseases, other polymorphisms found in a comprehensive range of genes need to be integrated to further comprehension.

Author Contributions

Conceptualization, M.B.-C.; writing—original draft preparation, C.Z., A.P.J.-U., J.P.-C. and M.B.-C.; writing—review and editing, C.Z., A.P.J.-U., J.P.-C. and M.B.-C.; figures preparation A.P.J.-U. and J.P.-C.; Table preparation, A.P.J.-U. and M.B.-C.; funding acquisition, M.B.-C. All authors have read and agreed to the published version of the manuscript.

Funding

Open access funding for this article was supported by Instituto Nacional de Cardiología, Ignacio Chavez (Mexico).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cardiometabolic Diseases. 2022. Available online: https://www.who.int/news-room/fact-sheets/detail/cardiovascular-diseases-(cvds) (accessed on 28 January 2022).
  2. Barteková, M.; Adameová, A.; Görbe, A.; Ferenczyová, K.; Pecháňová, O.; Lazou, A.; Dhalla, N.S.; Ferdinandy, P.; Giricz, Z. Natural and synthetic antioxidants targeting cardiac oxidative stress and redox signaling in cardiometabolic diseases. Free Radic. Biol. Med. 2021, 169, 446–477. [Google Scholar] [CrossRef] [PubMed]
  3. Patel, B.; Mann, G.E.; Chapple, S.J. Concerted redox modulation by sulforaphane alleviates diabetes and cardiometabolic syndrome. Free Radic. Biol. Med. 2018, 122, 150–160. [Google Scholar] [CrossRef] [PubMed]
  4. Yamamoto, T.; Yoh, K.; Kobayashi, A.; Ishii, Y.; Kure, S.; Koyama, A.; Sakamoto, T.; Sekizawa, K.; Motohashi, H.; Yamamoto, M. Identification of polymorphisms in the promoter region of the human NRF2 gene. Biochem. Biophys. Res. Commun. 2004, 321, 72–79. [Google Scholar] [CrossRef] [PubMed]
  5. Hua, C.-C.; Chang, L.-C.; Tseng, J.-C.; Chu, C.-M.; Liu, Y.-C.; Shieh, W.-B. Functional haplotypes in the promoter region of transcription factor Nrf2 in chronic obstructive pulmonary disease. Dis. Markers 2010, 28, 185–193. [Google Scholar] [CrossRef] [PubMed]
  6. Cho, H.-Y.; Marzec, J.; Kleeberger, S.R. Functional polymorphisms in Nrf2: Implications for human disease. Free Radic. Biol. Med. 2015, 88, 362–372. [Google Scholar] [CrossRef]
  7. Villalobos, J.B.; Molina-Muñoz, T.; Mailloux-Salinas, P.; Bravo, G.; Carvajal, K.; Gómez-Víquez, N.L. Oxidative stress in cardiomyocytes contributes to decreased SERCA2a activity in rats with metabolic syndrome. Am. J. Physiol. Heart Circ. Physiol. 2013, 305, H1344–H1353. [Google Scholar] [CrossRef] [Green Version]
  8. Ramírez-Camacho, I.; García-Niño, W.; Flores-García, M.; Pedraza-Chaverri, J.; Zazueta, C. Alteration of mitochondrial supercomplexes assembly in metabolic diseases. Biochim. Biophys. Acta (BBA)-Mol. Basis Dis. 2020, 1866, 165935. [Google Scholar] [CrossRef]
  9. Pisoschi, A.M.; Pop, A. The role of antioxidants in the chemistry of oxidative stress: A review. Eur. J. Med. Chem. 2015, 97, 55–74. [Google Scholar] [CrossRef]
  10. Lee, Y.; Im, E. Regulation of miRNAs by Natural Antioxidants in Cardiovascular Diseases: Focus on SIRT1 and eNOS. Antioxidants 2021, 10, 377. [Google Scholar] [CrossRef]
  11. Hayes, J.D.; Dinkova-Kostova, A. The Nrf2 regulatory network provides an interface between redox and intermediary metabolism. Trends Biochem. Sci. 2014, 39, 199–218. [Google Scholar] [CrossRef]
  12. Ibrahim, L.; Mesgarzadeh, J.; Xu, I.; Powers, E.T.; Wiseman, R.L.; Bollong, M.J. Defining the Functional Targets of Cap‘n’collar Transcription Factors NRF1, NRF2, and NRF3. Antioxidants 2020, 9, 1025. [Google Scholar] [CrossRef] [PubMed]
  13. Wakabayashi, N.; Slocum, S.L.; Skoko, J.J.; Shin, S.; Kensler, T.W. When NRF2 Talks, Who’s Listening? Antioxid. Redox Signal. 2010, 13, 1649–1663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. 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] [PubMed] [Green Version]
  15. Xue, Y.; Shui, X.; Su, W.; He, Y.; Lu, X.; Zhang, Y.; Yan, G.; Huang, S.; Lei, W.; Chen, C. Baicalin inhibits inflammation and attenuates myocardial ischaemic injury by aryl hydrocarbon receptor. J. Pharm. Pharmacol. 2015, 67, 1756–1764. [Google Scholar] [CrossRef] [PubMed]
  16. Silva-Palacios, A.; Ostolga-Chavarría, M.; Sánchez-Garibay, C.; Rojas-Morales, P.; Galván-Arzate, S.; Buelna-Chontal, M.; Pavón, N.; Pedraza-Chaverrí, J.; Königsberg, M.; Zazueta, C. Sulforaphane protects from myocardial ischemia-reperfusion damage through the balanced activation of Nrf2/AhR. Free Radic. Biol. Med. 2019, 143, 331–340. [Google Scholar] [CrossRef]
  17. Moi, P.; Chan, K.; Asunis, I.; Cao, A.; Kan, Y.W. Isolation of NF-E2-related factor 2 (Nrf2), a NF-E2-like basic leucine zipper transcriptional activator that binds to the tandem NF-E2/AP1 repeat of the beta-globin locus control region. Proc. Natl. Acad. Sci. USA 1994, 91, 9926–9930. [Google Scholar] [CrossRef] [Green Version]
  18. Itoh, K.; Wakabayashi, N.; Katoh, Y.; Ishii, T.; Igarashi, K.; Engel, J.D.; Yamamoto, M. Keap1 represses nuclear activation of antioxidant responsive elements by Nrf2 through binding to the amino-terminal Neh2 domain. Genes Dev. 1999, 13, 76–86. [Google Scholar] [CrossRef] [Green Version]
  19. Wang, H.; Liu, K.; Geng, M.; Gao, P.; Wu, X.; Hai, Y.; Li, Y.; Li, Y.; Luo, L.; Hayes, J.; et al. RXRα Inhibits the NRF2-ARE Signaling Pathway through a Direct Interaction with the Neh7 Domain of NRF2. Cancer Res. 2013, 73, 3097–3108. [Google Scholar] [CrossRef] [Green Version]
  20. Tong, K.I.; Katoh, Y.; Kusunoki, H.; Itoh, K.; Tanaka, T.; Yamamoto, M. Keap1 Recruits Neh2 through Binding to ETGE and DLG Motifs: Characterization of the Two-Site Molecular Recognition Model. Mol. Cell. Biol. 2006, 26, 2887–2900. [Google Scholar] [CrossRef] [Green Version]
  21. Katoh, Y.; Itoh, K.; Yoshida, E.; Miyagishi, M.; Fukamizu, A.; Yamamoto, M. Two domains of Nrf2 cooperatively bind CBP, a CREB binding protein, and synergistically activate transcription. Genes Cells 2001, 6, 857–868. [Google Scholar] [CrossRef]
  22. 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] [PubMed] [Green Version]
  23. Shen, G.; Hebbar, V.; Nair, S.; Xu, C.; Li, W.; Lin, W.; Keum, Y.-S.; Han, J.; Gallo, M.A.; Kong, A.-N.T. Regulation of Nrf2 Transactivation Domain Activity. The differential effects of mitogen-activated protein kinase cascades and synergistic stimulatory effect of Raf and CREB-binding protein. J. Biol. Chem. 2004, 279, 23052–23060. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Zhang, J.; Hosoya, T.; Maruyama, A.; Nishikawa, K.; Maher, J.M.; Ohta, T.; Motohashi, H.; Fukamizu, A.; Shibahara, S.; Itoh, K.; et al. Nrf2 Neh5 domain is differentially utilized in the transactivation of cytoprotective genes. Biochem. J. 2007, 404, 459–466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Li, W.; Yu, S.; Kong, A.-N.T. Nrf2 Possesses a Redox-sensitive Nuclear Exporting Signal in the Neh5 Transactivation Domain. J. Biol. Chem. 2006, 281, 27251–27263. [Google Scholar] [CrossRef] [Green Version]
  26. Chowdhry, S.; Zhang, Y.; McMahon, M.; Sutherland, C.; Cuadrado, A.; Hayes, J.D. Nrf2 is controlled by two distinct β-TrCP recognition motifs in its Neh6 domain, one of which can be modulated by GSK-3 activity. Oncogene 2013, 32, 3765–3781. [Google Scholar] [CrossRef] [Green Version]
  27. McMahon, M.; Thomas, N.; Itoh, K.; Yamamoto, M.; Hayes, J.D. Redox-regulated Turnover of Nrf2 Is Determined by at Least Two Separate Protein Domains, the Redox-sensitive Neh2 Degron and the Redox-insensitive Neh6 Degron. J. Biol. Chem. 2004, 279, 31556–31567. [Google Scholar] [CrossRef] [Green Version]
  28. Sun, Z.; Chin, Y.E.; Zhang, D.D. Acetylation of Nrf2 by p300/CBP Augments Promoter-Specific DNA Binding of Nrf2 during the Antioxidant Response. Mol. Cell. Biol. 2009, 29, 2658–2672. [Google Scholar] [CrossRef] [Green Version]
  29. Li, W.; Yu, S.; Liu, T.; Kim, J.-H.; Blank, V.; Li, H.; Kong, A.-N.T. Heterodimerization with small Maf proteins enhances nuclear retention of Nrf2 via masking the NESzip motif. Biochim. Biophys. Acta 2008, 1783, 1847–1856. [Google Scholar] [CrossRef] [Green Version]
  30. 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]
  31. Kobayashi, A.; Kang, M.-I.; Okawa, H.; Ohtsuji, M.; Zenke, Y.; Chiba, T.; Igarashi, K.; Yamamoto, M. Oxidative Stress Sensor Keap1 Functions as an Adaptor for Cul3-Based E3 Ligase To Regulate Proteasomal Degradation of Nrf2. Mol. Cell. Biol. 2004, 24, 7130–7139. [Google Scholar] [CrossRef] [Green Version]
  32. McMahon, M.; Thomas, N.; Itoh, K.; Yamamoto, M.; Hayes, J. Dimerization of Substrate Adaptors Can Facilitate Cullin-mediated Ubiquitylation of Proteins by a “Tethering” Mechanism: A two-site interaction model for the Nrf2-Keap1 complex. J. Biol. Chem. 2006, 281, 24756–24768. [Google Scholar] [CrossRef] [Green Version]
  33. Zhang, D.D.; Lo, S.-C.; Cross, J.V.; Templeton, D.J.; Hannink, M. Keap1 Is a Redox-Regulated Substrate Adaptor Protein for a Cul3-Dependent Ubiquitin Ligase Complex. Mol. Cell. Biol. 2004, 24, 10941–10953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Kaspar, J.W.; Jaiswal, A.K. An Autoregulatory Loop between Nrf2 and Cul3-Rbx1 Controls Their Cellular Abundance. J. Biol. Chem. 2010, 285, 21349–21358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Rada, P.; Rojo, A.I.; Chowdhry, S.; McMahon, M.; Hayes, J.D.; Cuadrado, A. SCF/-TrCP Promotes Glycogen Synthase Kinase 3-Dependent Degradation of the Nrf2 Transcription Factor in a Keap1-Independent Manner. Mol. Cell. Biol. 2011, 31, 1121–1133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Wang, L.; Wang, Y.; Zhang, C.; Li, J.; Meng, Y.; Dou, M.; Noguchi, C.T.; Di, L. Inhibiting Glycogen Synthase Kinase 3 Reverses Obesity-Induced White Adipose Tissue Inflammation by Regulating Apoptosis Inhibitor of Macrophage/CD5L-Mediated Macrophage Migration. Arter. Thromb. Vasc. Biol. 2018, 38, 2103–2116. [Google Scholar] [CrossRef] [PubMed]
  37. Eldar-Finkelman, H.; Schreyer, S.A.; Shinohara, M.M.; LeBoeuf, R.C.; Krebs, E.G. Increased glycogen synthase kinase-3 activity in diabetes- and obesity-prone C57BL/6J mice. Diabetes 1999, 48, 1662–1666. [Google Scholar] [CrossRef] [PubMed]
  38. Wu, W.; Liu, X.; Han, L. Apoptosis of cardiomyocytes in diabetic cardiomyopathy involves overexpression of glycogen synthase kinase-3β. Biosci. Rep. 2019, 39, BSR20171307. [Google Scholar] [CrossRef] [Green Version]
  39. 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] [Green Version]
  40. Takaya, K.; Suzuki, T.; Motohashi, H.; Onodera, K.; Satomi, S.; Kensler, T.W.; Yamamoto, M. Validation of the multiple sensor mechanism of the Keap1-Nrf2 system. Free Radic. Biol. Med. 2012, 53, 817–827. [Google Scholar] [CrossRef] [Green Version]
  41. McMahon, M.; Lamont, D.J.; Beattie, K.A.; Hayes, J.D. Keap1 perceives stress via three sensors for the endogenous signaling molecules nitric oxide, zinc, and alkenals. Proc. Natl. Acad. Sci. USA 2010, 107, 18838–18843. [Google Scholar] [CrossRef] [Green Version]
  42. Luo, Y.; Eggler, A.L.; Liu, D.; Liu, G.; Mesecar, A.D.; van Breemen, R.B. Sites of alkylation of human Keap1 by natural chemoprevention agents. J. Am. Soc. Mass Spectrom. 2007, 18, 2226–2232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. 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]
  44. Theodore, M.; Kawai, Y.; Yang, J.; Kleshchenko, Y.; Reddy, S.P.; Villalta, F.; Arinze, I.J. Multiple Nuclear Localization Signals Function in the Nuclear Import of the Transcription Factor Nrf2. J. Biol. Chem. 2008, 283, 8984–8994, Erratum in J. Biol. Chem. 2008, 283, 14176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Katsuoka, F.; Motohashi, H.; Ishii, T.; Aburatani, H.; Engel, J.D.; Yamamoto, M. Genetic Evidence that Small Maf Proteins Are Essential for the Activation of Antioxidant Response Element-Dependent Genes. Mol. Cell. Biol. 2005, 25, 8044–8051. [Google Scholar] [CrossRef] [Green Version]
  46. Bloom, D.A.; Jaiswal, A.K. Phosphorylation of Nrf2 at Ser40 by Protein Kinase C in Response to Antioxidants Leads to the Release of Nrf2 from INrf2, but Is Not Required for Nrf2 Stabilization/Accumulation in the Nucleus and Transcriptional Activation of Antioxidant Response Element-mediated NAD(P)H:Quinone Oxidoreductase-1 Gene Expression. J. Biol. Chem. 2003, 278, 44675–44682. [Google Scholar] [CrossRef] [Green Version]
  47. Apopa, P.L.; He, X.; Ma, Q. Phosphorylation of Nrf2 in the transcription activation domain by casein kinase 2 (CK2) is critical for the nuclear translocation and transcription activation function of Nrf2 in IMR-32 neuroblastoma cells. J. Biochem. Mol. Toxicol. 2008, 22, 63–76. [Google Scholar] [CrossRef]
  48. Jang, D.E.; Song, J.; Park, J.-W.; Yoon, S.-H.; Bae, Y.-S. Protein kinase CK2 activates Nrf2 via autophagic degradation of Keap1 and activation of AMPK in human cancer cells. BMB Rep. 2020, 53, 272–277. [Google Scholar] [CrossRef]
  49. Joo, M.S.; Kim, W.D.; Lee, K.Y.; Kim, J.H.; Koo, J.H.; Kim, S.G. AMPK Facilitates Nuclear Accumulation of Nrf2 by Phosphorylating at Serine 550. Mol. Cell. Biol. 2016, 36, 1931–1942. [Google Scholar] [CrossRef] [Green Version]
  50. Matzinger, M.; Fischhuber, K.; Pölöske, D.; Mechtler, K.; Heiss, E.H. AMPK leads to phosphorylation of the transcription factor Nrf2, tuning transactivation of selected target genes. Redox Biol. 2020, 29, 101393. [Google Scholar] [CrossRef]
  51. Volz, N.; Boettler, U.; Winkler, S.; Teller, N.; Schwarz, C.; Bakuradze, T.; Eisenbrand, G.; Haupt, L.M.; Griffiths, L.; Stiebitz, H.; et al. Effect of Coffee Combining Green Coffee Bean Constituents with Typical Roasting Products on the Nrf2/ARE Pathway in Vitro and in Vivo. J. Agric. Food Chem. 2012, 60, 9631–9641. [Google Scholar] [CrossRef]
  52. Boettler, U.; Volz, N.; Teller, N.; Haupt, L.M.; Bakuradze, T.; Eisenbrand, G.; Bytof, G.; Lantz, I.; Griffiths, L.R.; Marko, D. Induction of antioxidative Nrf2 gene transcription by coffee in humans: Depending on genotype? Mol. Biol. Rep. 2012, 39, 7155–7162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Jerotic, D.; Matic, M.; Suvakov, S.; Vucicevic, K.; Damjanovic, T.; Savic-Radojevic, A.; Pljesa-Ercegovac, M.; Coric, V.; Stefanovic, A.; Ivanisevic, J.; et al. Association of Nrf2, SOD2 and GPX1 Polymorphisms with Biomarkers of Oxidative Distress and Survival in End-Stage Renal Disease Patients. Toxins 2019, 11, 431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Dos Santos, K.N.; Florentino, R.M.; França, A.; Filho, A.C.M.L.; Dos Santos, M.L.; Missiaggia, D.; Fonseca, M.D.C.; Costa, I.B.; Vidigal, P.V.T.; Nathanson, M.H.; et al. Polymorphism in the Promoter Region of NFE2L2 Gene Is a Genetic Marker of Susceptibility to Cirrhosis Associated with Alcohol Abuse. Int. J. Mol. Sci. 2019, 20, 3589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Korytina, G.F.; Akhmadishina, L.Z.; Aznabaeva, Y.; Kochetova, O.V.; Zagidullin, N.; Kzhyshkowska, J.G.; Zagidullin, S.Z.; Viktorova, T.V. Associations of the NRF2/KEAP1 pathway and antioxidant defense gene polymorphisms with chronic obstructive pulmonary disease. Gene 2019, 692, 102–112. [Google Scholar] [CrossRef] [PubMed]
  56. Siedlinski, M.; Postma, D.S.; Boer, J.M.; van der Steege, G.; Schouten, J.P.; Smit, H.A.; Boezen, H.M. Level and course of FEV1 in relation to polymorphisms in NFE2L2 and KEAP1 in the general population. Respir. Res. 2009, 10, 73. [Google Scholar] [CrossRef] [Green Version]
  57. Testa, A.; Leonardis, D.; Spoto, B.; Sanguedolce, M.C.; Parlongo, R.M.; Pisano, A.; Tripepi, G.; Mallamaci, F.; Zoccali, C. A polymorphism in a major antioxidant gene (Kelch-like ECH-associated protein 1) predicts incident cardiovascular events in chronic kidney disease patients: An exploratory study. J. Hypertens. 2016, 34, 928–934. [Google Scholar] [CrossRef]
  58. Marzec, J.M.; Christie, J.D.; Reddy, S.P.; Jedlicka, A.E.; Vuong, H.; Lanken, P.N.; Aplenc, R.; Yamamoto, T.; Yamamoto, M.; Cho, H.-Y.; et al. Functional polymorphisms in the transcription factor NRF2 in humans increase the risk of acute lung injury. FASEB J. 2007, 21, 2237–2246. [Google Scholar] [CrossRef]
  59. Figarska, S.M.; Vonk, J.M.; Boezen, H.M. NFE2L2 polymorphisms, mortality, and metabolism in the general population. Physiol. Genom. 2014, 46, 411–417. [Google Scholar] [CrossRef]
  60. Scutt, G.; Overall, A.; Bakrania, P.; Krasteva, E.; Parekh, N.; Ali, K.; Davies, J.G.; Rajkumar, C. The Association of a Single-Nucleotide Polymorphism in the Nuclear Factor (Erythroid-Derived 2)-Like 2 Gene With Adverse Drug Reactions, Multimorbidity, and Frailty in Older People. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 1050–1057. [Google Scholar] [CrossRef]
  61. Rohm, T.V.; Meier, D.T.; Olefsky, J.M.; Donath, M.Y. Inflammation in obesity, diabetes, and related disorders. Immunity 2022, 55, 31–55. [Google Scholar] [CrossRef]
  62. Ozata, M.; Mergen, M.; Oktenli, C.; Aydin, A.; Sanisoglu, S.Y.; Bolu, E.; Yilmaz, M.; Sayal, A.; Isimer, A.; Ozdemir, I. Increased oxidative stress and hypozincemia in male obesity. Clin. Biochem. 2002, 35, 627–631. [Google Scholar] [CrossRef]
  63. Adnan, T.; Amin, M.N.; Uddin, G.; Hussain, S.; Sarwar, S.; Hossain, K.; Uddin, S.N.; Islam, M.S. Increased concentration of serum MDA, decreased antioxidants and altered trace elements and macro-minerals are linked to obesity among Bangladeshi population. Diabetes Metab. Syndr. Clin. Res. Rev. 2019, 13, 933–938. [Google Scholar] [CrossRef] [PubMed]
  64. Monzo-Beltran, L.; Vazquez-Tarragón, A.; Cerdà, C.; Garcia-Perez, P.; Iradi, A.; Sánchez, C.; Climent, B.; Tormos, C.; Vázquez-Prado, A.; Girbés, J.; et al. One-year follow-up of clinical, metabolic and oxidative stress profile of morbid obese patients after laparoscopic sleeve gastrectomy. 8-oxo-dG as a clinical marker. Redox Biol. 2017, 12, 389–402. [Google Scholar] [CrossRef] [PubMed]
  65. Jankovic, A.; Korac, A.; Galic, B.S.; Buzadzic, B.; Otasevic, V.; Stancic, A.; Vucetic, M.; Markelić, M.; Velickovic, K.; Golic, I.; et al. Differences in the redox status of human visceral and subcutaneous adipose tissues—Relationships to obesity and metabolic risk. Metabolism 2014, 63, 661–671. [Google Scholar] [CrossRef] [PubMed]
  66. Malik, A.; Morya, R.K.; Saha, S.; Singh, P.K.; Bhadada, S.K.; Rana, S.V. Oxidative stress and inflammatory markers in type 2 diabetic patients. Eur. J. Clin. Investig. 2020, 50, e13238. [Google Scholar] [CrossRef]
  67. Nowotny, K.; Jung, T.; Höhn, A.; Weber, D.; Grune, T.; Nowotny, K.; Jung, T.; Höhn, A.; Weber, D.; Grune, T. Advanced Glycation End Products and Oxidative Stress in Type 2 Diabetes Mellitus. Biomolecules 2015, 5, 194–222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Li, W.; Zhu, C.; Liu, T.; Zhang, W.; Liu, X.; Li, P.; Zhu, T. Epigallocatechin-3-gallate ameliorates glucolipid metabolism and oxidative stress in type 2 diabetic rats. Diabetes Vasc. Dis. Res. 2020, 17, 1479164120966998. [Google Scholar] [CrossRef]
  69. Wang, W.; He, Y.; Liu, Q. Parthenolide plays a protective role in the liver of mice with metabolic dysfunction-associated fatty liver disease through the activation of the HIPPO pathway. Mol. Med. Rep. 2021, 24, 487. [Google Scholar] [CrossRef]
  70. Perdomo, L.; Beneit, N.; Otero, Y.F.; Escribano, Ó.; Díaz-Castroverde, S.; Gómez-Hernández, A.; Benito, M. Protective role of oleic acid against cardiovascular insulin resistance and in the early and late cellular atherosclerotic process. Cardiovasc. Diabetol. 2015, 14, 75. [Google Scholar] [CrossRef] [Green Version]
  71. Zurbau, A.; Duvnjak, L.S.; Magas, S.; Jovanovski, E.; Miocic, J.; Jenkins, A.L.; Jenkins, D.J.A.; Josse, R.G.; Leiter, L.A.; Sievenpiper, J.L.; et al. Co-administration of viscous fiber, Salba-chia and ginseng on glycemic management in type 2 diabetes: A double-blind randomized controlled trial. Eur. J. Nutr. 2021, 60, 3071–3083. [Google Scholar] [CrossRef]
  72. Ren, B.C.; Zhang, Y.F.; Liu, S.S.; Cheng, X.J.; Yang, X.; Cui, X.G.; Zhao, X.R.; Zhao, H.; Hao, M.F.; Li, M.D.; et al. Curcumin alleviates oxidative stress and inhibits apoptosis in diabetic cardiomyopathy via Sirt1-Foxo1 and PI3K-Akt signalling pathways. J. Cell. Mol. Med. 2020, 24, 12355–12367. [Google Scholar] [CrossRef] [PubMed]
  73. Tian, S.; Wang, Y.; Li, X.; Liu, J.; Wang, J.; Lu, Y. Sulforaphane Regulates Glucose and Lipid Metabolisms in Obese Mice by Restraining JNK and Activating Insulin and FGF21 Signal Pathways. J. Agric. Food Chem. 2021, 69, 13066–13079. [Google Scholar] [CrossRef] [PubMed]
  74. Wang, P.; Gao, J.; Ke, W.; Wang, J.; Li, D.; Liu, R.; Jia, Y.; Wang, X.; Chen, X.; Chen, F.; et al. Resveratrol reduces obesity in high-fat diet-fed mice via modulating the composition and metabolic function of the gut microbiota. Free Radic. Biol. Med. 2020, 156, 83–98. [Google Scholar] [CrossRef] [PubMed]
  75. Su, X.; Wang, S.; Zhang, H.; Yang, G.; Bai, Y.; Liu, P.; Meng, L.; Jiang, X.; Xin, Y. Sulforaphane prevents angiotensin II-induced cardiomyopathy by activation of Nrf2 through epigenetic modification. J. Cell. Mol. Med. 2021, 25, 4408–4419. [Google Scholar] [CrossRef] [PubMed]
  76. Darband, S.G.; Sadighparvar, S.; Yousefi, B.; Kaviani, M.; Ghaderi-Pakdel, F.; Mihanfar, A.; Rahimi, Y.; Mobaraki, K.; Majidinia, M. Quercetin attenuated oxidative DNA damage through NRF2 signaling pathway in rats with DMH induced colon carcinogenesis. Life Sci. 2020, 253, 117584. [Google Scholar] [CrossRef]
  77. Zhao, P.; Hu, Z.; Ma, W.; Zang, L.; Tian, Z.; Hou, Q. Quercetin alleviates hyperthyroidism-induced liver damage via Nrf2 Signaling pathway. BioFactors 2020, 46, 608–619. [Google Scholar] [CrossRef]
  78. Tan, Y.; Ichikawa, T.; Li, J.; Si, Q.; Yang, H.; Chen, X.; Goldblatt, C.S.; Meyer, C.J.; Li, X.; Cai, L.; et al. Diabetic Downregulation of Nrf2 Activity via ERK Contributes to Oxidative Stress–Induced Insulin Resistance in Cardiac Cells In Vitro and In Vivo. Diabetes 2011, 60, 625–633. [Google Scholar] [CrossRef] [Green Version]
  79. Uruno, A.; Yagishita, Y.; Yamamoto, M. The Keap1–Nrf2 system and diabetes mellitus. Arch. Biochem. Biophys. 2015, 566, 76–84. [Google Scholar] [CrossRef] [Green Version]
  80. Adam, M.; Imboden, M.; Schaffner, E.; Boes, E.; Kronenberg, F.; Pons, M.; Bettschart, R.; Barthelemy, J.-C.; Schindler, C.; Probst-Hensch, N. The adverse impact of obesity on heart rate variability is modified by a NFE2L2 gene variant: The SAPALDIA cohort. Int. J. Cardiol. 2017, 228, 341–346. [Google Scholar] [CrossRef] [Green Version]
  81. Gusti, A.; Qusti, S.; Alshammari, E.; Toraih, E.; Fawzy, M. Antioxidants-Related Superoxide Dismutase (SOD), Catalase (CAT), Glutathione Peroxidase (GPX), Glutathione-S-Transferase (GST), and Nitric Oxide Synthase (NOS) Gene Variants Analysis in an Obese Population: A Preliminary Case-Control Study. Antioxidants 2021, 10, 595. [Google Scholar] [CrossRef]
  82. Montano, M.A.E.; Lera, J.P.B.; Gottlieb, M.; Schwanke, C.; Da Rocha, M.I.U.M.; Manica-Cattani, M.F.; Dos Santos, G.F.; da Cruz, I. Association between manganese superoxide dismutase (MnSOD) gene polymorphism and elderly obesity. Mol. Cell. Biochem. 2009, 328, 33–40. [Google Scholar] [CrossRef] [PubMed]
  83. Duarte, T.; da Cruz, I.; Barbisan, F.; Capelleto, D.; Moresco, R.; Duarte, M.M.M.F. The effects of rosuvastatin on lipid-lowering, inflammatory, antioxidant and fibrinolytics blood biomarkers are influenced by Val16Ala superoxide dismutase manganese-dependent gene polymorphism. Pharm. J. 2016, 16, 501–506. [Google Scholar] [CrossRef] [PubMed]
  84. Gottlieb, M.G.V.; Schwanke, C.H.A.; Santos, A.F.R.; Jobim, P.F.; Müssel, D.P.; da Cruz, I. Association among oxidized LDL levels, MnSOD, apolipoprotein E polymorphisms, and cardiovascular risk factors in a south Brazilian region population. Genet. Mol. Res. 2005, 4, 691–703. [Google Scholar] [PubMed]
  85. Nakanishi, S.; Yamane, K.; Ohishi, W.; Nakashima, R.; Yoneda, M.; Nojima, H.; Watanabe, H.; Kohno, N. Manganese superoxide dismutase Ala16Val polymorphism is associated with the development of type 2 diabetes in Japanese-Americans. Diabetes Res. Clin. Pract. 2008, 81, 381–385. [Google Scholar] [CrossRef] [PubMed]
  86. Nomiyama, T.; Tanaka, Y.; Piao, L.; Nagasaka, K.; Sakai, K.; Ogihara, T.; Nakajima, K.; Watada, H.; Kawamori, R. The polymorphism of manganese superoxide dismutase is associated with diabetic nephropathy in Japanese type 2 diabetic patients. J. Hum. Genet. 2003, 48, 0138–0141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Mollsten, A.; Marklund, S.L.; Wessman, M.; Svensson, M.; Forsblom, C.; Parkkonen, M.; Brismar, K.; Groop, P.-H.; Dahlquist, G. A Functional Polymorphism in the Manganese Superoxide Dismutase Gene and Diabetic Nephropathy. Diabetes 2007, 56, 265–269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Mollsten, A.; Jorsal, A.; Lajer, M.; Vionnet, N.; Tarnow, L. The V16A polymorphism in SOD2 is associated with increased risk of diabetic nephropathy and cardiovascular disease in type 1 diabetes. Diabetologia 2009, 52, 2590–2593. [Google Scholar] [CrossRef] [Green Version]
  89. Xu, X.; Sun, J.; Chang, X.; Wang, J.; Luo, M.; Wintergerst, K.A.; Miao, L.; Cai, L. Genetic variants of nuclear factor erythroid-derived 2-like 2 associated with the complications in Han descents with type 2 diabetes mellitus of Northeast China. J. Cell. Mol. Med. 2016, 20, 2078–2088. [Google Scholar] [CrossRef] [Green Version]
  90. Hernández Guerrero, C.; Hernández Chávez, P.; Martínez Castro, N.; Parra Carriedo, A.; García Del Rio, S.; Pérez Lizaur, A. Glutathione peroxidase-1 PRO200LEU polymorphism (Rs1050450) is associated with morbid obesity independently of the presence of prediabetes or diabetes in women from central Mexico. Nutr. Hosp. 2015, 32, 1516–1525. [Google Scholar] [CrossRef]
  91. Buraczynska, M.; Buraczynska, K.; Dragan, M.; Ksiazek, A. Pro198Leu Polymorphism in the Glutathione Peroxidase 1 Gene Contributes to Diabetic Peripheral Neuropathy in Type 2 Diabetes Patients. Neuromolecular Med. 2017, 19, 147–153. [Google Scholar] [CrossRef] [Green Version]
  92. Tang, T.; Prior, S.; Li, K.; Ireland, H.; Bain, S.; Hurel, S.; Cooper, J.; Humphries, S.; Stephens, J. Association between the rs1050450 glutathione peroxidase-1 (C > T) gene variant and peripheral neuropathy in two independent samples of subjects with diabetes mellitus. Nutr. Metab. Cardiovasc. Dis. 2012, 22, 417–425. [Google Scholar] [CrossRef]
  93. Liu, D.; Liu, L.; Hu, Z.; Song, Z.; Wang, Y.; Chen, Z. Evaluation of the oxidative stress–related genes ALOX5, ALOX5AP, GPX1, GPX3 and MPO for contribution to the risk of type 2 diabetes mellitus in the Han Chinese population. Diabetes Vasc. Dis. Res. 2018, 15, 336–339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Jiménez-Osorio, A.S.; González-Reyes, S.; García-Niño, W.R.; Moreno-Macías, H.; Arellano, M.E.R.; Vargas-Alarcón, G.; Zuñiga, J.; Barquera, R.; Pedraza-Chaverri, J. Association of Nuclear Factor-Erythroid 2-Related Factor 2, Thioredoxin Interacting Protein, and Heme Oxygenase-1 Gene Polymorphisms with Diabetes and Obesity in Mexican Patients. Oxidative Med. Cell. Longev. 2016, 2016, 7367641, Erratum in Oxidative Med. Cell. Longev. 2017, 2017, 7543194. [Google Scholar] [CrossRef] [PubMed]
  95. Wang, J.; Yu, M.; Chen, J.; Zhu, L.; Liu, J.; Xu, J. Association of Nuclear Factor Erythroid-2-Related Actor 2 Gene Polymorphisms with Diabetic Nephropathy in Chinese Patients. Int. J. Gen. Med. 2021, 14, 1231–1237. [Google Scholar] [CrossRef] [PubMed]
  96. Wang, X.; Chen, H.; Liu, J.; Ouyang, Y.; Wang, D.; Bao, W.; Liu, L. Association between the NF-E2 Related Factor 2 Gene Polymorphism and Oxidative Stress, Anti-Oxidative Status, and Newly-Diagnosed Type 2 Diabetes Mellitus in a Chinese Population. Int. J. Mol. Sci. 2015, 16, 16483–16496. [Google Scholar] [CrossRef] [Green Version]
  97. Teena, R.; Dhamodharan, U.; Ali, D.; Rajesh, K.; Ramkumar, K.M. Genetic Polymorphism of the Nrf2 Promoter Region (rs35652124) Is Associated with the Risk of Diabetic Foot Ulcers. Oxidative Med. Cell. Longev. 2020, 2020, 9825028. [Google Scholar] [CrossRef]
  98. Teena, R.; Dhamodharan, U.; Jayasuriya, R.; Ali, D.; Kesavan, R.; Ramkumar, K.M. Analysis of the Exonic Single Nucleotide Polymorphism rs182428269 of the NRF2 Gene in Patients with Diabetic Foot Ulcer. Arch. Med. Res. 2021, 52, 224–232. [Google Scholar] [CrossRef]
  99. Shin, J.H.; Lee, K.-M.; Shin, J.; Kang, K.D.; Nho, C.W.; Cho, Y.S. Genetic risk score combining six genetic variants associated with the cellular NRF2 expression levels correlates with Type 2 diabetes in the human population. Genes Genom. 2019, 41, 537–545. [Google Scholar] [CrossRef]
  100. Da Costa, R.M.; Rodrigues, D.; Pereira, C.A.; Silva, J.; Alves, J.V.; Lobato, N.S.; Tostes, R.C. Nrf2 as a Potential Mediator of Cardiovascular Risk in Metabolic Diseases. Front. Pharmacol. 2019, 10, 382. [Google Scholar] [CrossRef] [Green Version]
  101. Howden, R. Nrf2and Cardiovascular Defense. Oxidative Med. Cell. Longev. 2013, 2013, 104308. [Google Scholar] [CrossRef] [Green Version]
  102. Buelna-Chontal, M.; Zazueta, C. Redox activation of Nrf2 & NF-κB: A double end sword? Cell. Signal. 2013, 25, 2548–2557. [Google Scholar] [CrossRef]
  103. 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]
  104. Shimoyama, Y.; Mitsuda, Y.; Tsuruta, Y.; Hamajima, N.; Niwa, T. Polymorphism of Nrf2, an Antioxidative Gene, is Associated with Blood Pressure and Cardiovascular Mortality in Hemodialysis Patients. Int. J. Med. Sci. 2014, 11, 726–731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Barajas, B.; Che, N.; Yin, F.; Rowshanrad, A.; Orozco, L.D.; Gong, K.W.; Wang, X.; Castellani, L.W.; Reue, K.; Lusis, A.J.; et al. NF-E2–Related Factor 2 Promotes Atherosclerosis by Effects on Plasma Lipoproteins and Cholesterol Transport That Overshadow Antioxidant Protection. Arter. Thromb. Vasc. Biol. 2011, 31, 58–66. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Freigang, S.; Ampenberger, F.; Spohn, G.; Heer, S.; Shamshiev, A.T.; Kisielow, J.; Hersberger, M.; Yamamoto, M.; Bachmann, M.F.; Kopf, M. Nrf2 is essential for cholesterol crystal-induced inflammasome activation and exacerbation of atherosclerosis. Eur. J. Immunol. 2011, 41, 2040–2051. [Google Scholar] [CrossRef] [PubMed]
  107. Ruotsalainen, A.-K.; Lappalainen, J.P.; Heiskanen, E.; Merentie, M.; Sihvola, V.; Näpänkangas, J.; Lottonen-Raikaslehto, L.; Kansanen, E.; Adinolfi, S.; Kaarniranta, K.; et al. Nuclear factor E2-related factor 2 deficiency impairs atherosclerotic lesion development but promotes features of plaque instability in hypercholesterolaemic mice. Cardiovasc. Res. 2018, 115, 243–254. [Google Scholar] [CrossRef] [PubMed]
  108. Sarutipaiboon, I.; Settasatian, N.; Komanasin, N.; Kukongwiriyapan, U.; Sawanyawisuth, K.; Intharaphet, P.; Senthong, V.; Settasatian, C. Association of Genetic Variations in NRF2, NQO1, HMOX1, and MT with Severity of Coronary Artery Disease and Related Risk Factors. Cardiovasc. Toxicol. 2020, 20, 176–189. [Google Scholar] [CrossRef]
  109. Miyazawa, M.; Tsuji, Y. Evidence for a novel antioxidant function and isoform-specific regulation of the human p66Shc gene. Mol. Biol. Cell 2014, 25, 2116–2127. [Google Scholar] [CrossRef]
  110. Napoli, C.; Martin-Padura, I.; de Nigris, F.; Giorgio, M.; Mansueto, G.; Somma, P.; Condorelli, M.; Sica, G.; de Rosa, G.; Pelicci, P. Deletion of the p66Shc longevity gene reduces systemic and tissue oxidative stress, vascular cell apoptosis, and early atherogenesis in mice fed a high-fat diet. Proc. Natl. Acad. Sci. USA 2003, 100, 2112–2116. [Google Scholar] [CrossRef] [Green Version]
  111. Sentinelli, F.; Romeo, S.; Barbetti, F.; Berni, A.; Filippi, E.; Fanelli, M.; Fallarino, M.; Baroni, M.G. Search for genetic variants in the p66Shc longevity gene by PCR-single strand conformational polymorphism in patients with early-onset cardiovascular disease. BMC Genet. 2006, 7, 14. [Google Scholar] [CrossRef] [Green Version]
  112. Raghunath, A.; Sundarraj, K.; Nagarajan, R.; Arfuso, F.; Bian, J.; Kumar, A.P.; Sethi, G.; Perumal, E. Antioxidant response elements: Discovery, classes, regulation and potential applications. Redox Biol. 2018, 17, 297–314. [Google Scholar] [CrossRef] [PubMed]
  113. Wang, X.; Campbell, M.R.; Lacher, S.E.; Cho, H.-Y.; Wan, M.; Crowl, C.L.; Chorley, B.N.; Bond, G.L.; Kleeberger, S.R.; Slattery, M.; et al. A Polymorphic Antioxidant Response Element Links NRF2/sMAF Binding to Enhanced MAPT Expression and Reduced Risk of Parkinsonian Disorders. Cell Rep. 2016, 15, 830–842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Linna-Kuosmanen, S.; Viitala, S.; Laitinen, T.; Peräkylä, M.; Pölönen, P.; Kansanen, E.; Leinonen, H.; Raju, S.; Wienecke-Baldacchino, A.; Närvänen, A.; et al. The Effects of Sequence Variation on Genome-wide NRF2 Binding—New Target Genes and Regulatory SNPs. Nucleic Acids Res. 2016, 44, 1760–1775. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Nishizawa, H.; Matsumoto, M.; Shindo, T.; Saigusa, D.; Kato, H.; Suzuki, K.; Sato, M.; Ishii, Y.; Shimokawa, H.; Igarashi, K. Ferroptosis is controlled by the coordinated transcriptional regulation of glutathione and labile iron metabolism by the transcription factor BACH1. J. Biol. Chem. 2020, 295, 69–82. [Google Scholar] [CrossRef] [PubMed]
  116. Ono, K.; Mannami, T.; Iwai, N. Association of a promoter variant of the haeme oxygenase-1 gene with hypertension in women. J. Hypertens. 2003, 21, 1497–1503. [Google Scholar] [CrossRef] [PubMed]
  117. Han, S.J.; Kang, E.S.; Kim, H.J.; Kim, S.H.; Chun, S.W.; Ahn, C.W.; Cha, B.S.; Nam, M.; Lee, H.C. The C609T variant of NQO1 is associated with carotid artery plaques in patients with type 2 diabetes. Mol. Genet. Metab. 2009, 97, 85–90. [Google Scholar] [CrossRef]
  118. Martin, N.J.; Collier, A.; Bowen, L.D.; Pritsos, K.L.; Goodrich, G.G.; Arger, K.; Cutter, G.; Pritsos, C.A. Polymorphisms in the NQO1, GSTT and GSTM genes are associated with coronary heart disease and biomarkers of oxidative stress. Mutat. Res. 2009, 674, 93–100. [Google Scholar] [CrossRef]
  119. Yu, H.; Liu, H.; Wang, L.-E.; Wei, Q. A Functional NQO1 609C>T Polymorphism and Risk of Gastrointestinal Cancers: A Meta-Analysis. PLoS ONE 2012, 7, e30566. [Google Scholar] [CrossRef] [Green Version]
  120. Duarte, M.M.; Moresco, R.N.; Duarte, T.; Santi, A.; Bagatini, M.D.; Da Cruz, I.B.; Schetinger, M.R.; Loro, V.L. Oxidative stress in hypercholesterolemia and its association with Ala16Val superoxide dismutase gene polymorphism. Clin. Biochem. 2010, 43, 1118–1123. [Google Scholar] [CrossRef]
  121. Lopes, R.A.; Neves, K.B.; Tostes, R.C.; Montezano, A.C.; Touyz, R.M. Downregulation of Nuclear Factor Erythroid 2–Related Factor and Associated Antioxidant Genes Contributes to Redox-Sensitive Vascular Dysfunction in Hypertension. Hypertension 2015, 66, 1240–1250. [Google Scholar] [CrossRef] [Green Version]
  122. Kunnas, T.; Määttä, K.; Nikkari, S.T. Genetic Polymorphisms of Transcription Factor NRF2 and of its Host Gene Sulfiredoxin (SRXN1) are Associated with Cerebrovascular Disease in a Finnish Cohort, the TAMRISK Study. Int. J. Med. Sci. 2016, 13, 325–329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Marczak, E.D.; Marzec, J.; Zeldin, D.; Kleeberger, S.; Brown, N.J.; Pretorius, M.; Lee, C.R. Polymorphisms in the transcription factor NRF2 and forearm vasodilator responses in humans. Pharm. Genom. 2012, 22, 620–628. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Ma, L.-L.; Sun, L.; Wang, Y.-X.; Sun, B.-H.; Li, Y.-F.; Jin, Y.-L. Association between HO-1 gene promoter polymorphisms and diseases (Review). Mol. Med. Rep. 2022, 25, 29. [Google Scholar] [CrossRef] [PubMed]
  125. Mansego, M.L.; Solar, G.D.M.; Alonso, M.P.; Martínez, F.; Sáez, G.T.; Escudero, J.C.M.; Redón, J.; Chaves, F.J. Polymorphisms of antioxidant enzymes, blood pressure and risk of hypertension. J. Hypertens. 2011, 29, 492–500. [Google Scholar] [CrossRef]
  126. An, S.H.; Chang, B.C.; Lee, K.E.; Gwak, H.S. Influence of UDP-Glucuronosyltransferase Polymorphisms on Stable Warfarin Doses in Patients with Mechanical Cardiac Valves. Cardiovasc. Ther. 2015, 33, 324–328. [Google Scholar] [CrossRef] [Green Version]
  127. Nakamura, A.; Nakajima, M.; Higashi, E.; Yamanaka, H.; Yokoi, T. Genetic polymorphisms in the 5′-flanking region of human UDP-glucuronosyltransferase 2B7 affect the Nrf2-dependent transcriptional regulation. Pharm. Genom. 2008, 18, 709–720. [Google Scholar] [CrossRef]
  128. Pereira, N.L.; Rihal, C.S.; So, D.Y.; Rosenberg, Y.; Lennon, R.J.; Mathew, V.; Goodman, S.G.; Weinshilboum, R.M.; Wang, L.; Baudhuin, L.M.; et al. Clopidogrel Pharmacogenetics. Circ. Cardiovasc. Interv. 2019, 12, e007811. [Google Scholar] [CrossRef]
  129. Pare, G.; Ross, S.; Mehta, S.R.; Yusuf, S.; Anand, S.S.; Connolly, S.J.; Fox, K.; Eikelboom, J.W. Effect of PON1 Q192R Genetic Polymorphism on Clopidogrel Efficacy and Cardiovascular Events in the Clopidogrel in the Unstable Angina to Prevent Recurrent Events Trial and the Atrial Fibrillation Clopidogrel Trial With Irbesartan for Prevention of Vascular Events. Circ. Cardiovasc. Genet. 2012, 5, 250–256. [Google Scholar] [CrossRef] [Green Version]
  130. Kosmidou, M.; Hatzitolios, A.I.; Molyva, D.; Raikos, N.; Savopoulos, C.; Daferera, N.; Kokkas, V.; Goulas, A. An association study between catalase -262C>T gene polymorphism, sodium-lithium countertrasport activity, insulin resistance, blood lipid parameters and their response to atorvastatin, in Greek dyslipidaemic patients and normolipidaemic controls. Free Radic. Res. 2009, 43, 385–389. [Google Scholar] [CrossRef]
  131. Watanabe, L.M.; Bueno, A.C.; de Lima, L.F.; Ferraz-Bannitz, R.; Dessordi, R.; Guimarães, M.P.; Foss-Freitas, M.C.; Barbosa, F., Jr.; Navarro, A.M. Genetically determined variations of selenoprotein P are associated with antioxidant, muscular, and lipid biomarkers in response to Brazil nut consumption by patients using statins. Br. J. Nutr. 2021, 127, 679–686. [Google Scholar] [CrossRef]
  132. Van Schaik, L.; Kettle, C.; Green, R.; Irving, H.R.; Rathner, J.A. Effects of Caffeine on Brown Adipose Tissue Thermogenesis and Metabolic Homeostasis: A Review. Front. Neurosci. 2021, 15, 621356. [Google Scholar] [CrossRef]
  133. Lara-Guzmán, O.J.; Álvarez, R.; Muñoz-Durango, K. Changes in the plasma lipidome of healthy subjects after coffee consumption reveal potential cardiovascular benefits: A randomized controlled trial. Free Radic. Biol. Med. 2021, 176, 345–355. [Google Scholar] [CrossRef] [PubMed]
  134. Krittanawong, C.; Tunhasiriwet, A.; Wang, Z.; Farrell, A.M.; Chirapongsathorn, S.; Zhang, H.; Kitai, T.; Mehta, D. Is caffeine or coffee consumption a risk for new-onset atrial fibrillation? A systematic review and meta-analysis. Eur. J. Prev. Cardiol. 2021, 28, e13–e15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Kolb, H.; Martin, S.; Kempf, K. Coffee and Lower Risk of Type 2 Diabetes: Arguments for a Causal Relationship. Nutrients 2021, 13, 1144. [Google Scholar] [CrossRef] [PubMed]
  136. Ye, X.; Liu, Y.; Hu, J.; Gao, Y.; Ma, Y.; Wen, D. Chlorogenic Acid-Induced Gut Microbiota Improves Metabolic Endotoxemia. Front. Endocrinol. 2021, 12, 762691. [Google Scholar] [CrossRef]
  137. Redondo-Puente, M.; Mateos, R.; Seguido, M.A.; García-Cordero, J.; González, S.; Tarradas, R.M.; Bravo-Clemente, L.; Sarriá, B. Appetite and Satiety Effects of the Acute and Regular Consumption of Green Coffee Phenols and Green Coffee Phenol/Oat β-Glucan Nutraceuticals in Subjects with Overweight and Obesity. Foods 2021, 10, 2511. [Google Scholar] [CrossRef]
  138. Mateos, R.; García-Cordero, J.; Bravo-Clemente, L.; Sarriá, B. Evaluation of novel nutraceuticals based on the combination of oat beta-glucans and a green coffee phenolic extract to combat obesity and its comorbidities. A randomized, dose–response, parallel trial. Food Funct. 2021, 13, 574–586. [Google Scholar] [CrossRef]
  139. Stevens, L.M.; Linstead, E.; Hall, J.L.; Kao, D.P. Association Between Coffee Intake and Incident Heart Failure Risk: A Machine Learning Analysis of the FHS, the ARIC Study, and the CHS. Circ. Heart Fail. 2021, 14, e006799. [Google Scholar] [CrossRef]
  140. Zhou, A.; Hyppönen, E. Habitual coffee intake and plasma lipid profile: Evidence from UK Biobank. Clin. Nutr. 2021, 40, 4404–4413. [Google Scholar] [CrossRef]
  141. Zhang, Z.; Wang, M.; Yuan, S.; Liu, X. Coffee consumption and risk of coronary artery disease. Eur. J. Prev. Cardiol. 2020, 29, e29–e31. [Google Scholar] [CrossRef]
  142. Nordestgaard, A.T. Causal relationship from coffee consumption to diseases and mortality: A review of observational and Mendelian randomization studies including cardiometabolic diseases, cancer, gallstones and other diseases. Eur. J. Nutr. 2021, 61, 573–587. [Google Scholar] [CrossRef] [PubMed]
  143. Siller-Lopez, F.; Garzón-Castaño, S.; Ramos-Márquez, M.E.; Hernandez-Cañaveral, I. Association of Paraoxonase-1 Q192R (rs662) Single Nucleotide Variation with Cardiovascular Risk in Coffee Harvesters of Central Colombia. J. Toxicol. 2017, 2017, 6913106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Bayram, B.; Ozcelik, B.; Grimm, S.; Roeder, T.; Schrader, C.; Ernst, I.M.; Wagner, A.E.; Grune, T.; Frank, J.; Rimbach, G. A Diet Rich in Olive Oil Phenolics Reduces Oxidative Stress in the Heart of SAMP8 Mice by Induction of Nrf2-Dependent Gene Expression. Rejuvenation Res. 2012, 15, 71–81. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Hassmann, U.; Haupt, L.M.; Smith, R.A.; Winkler, S.; Bytof, G.; Lantz, I.; Griffiths, L.R.; Marko, D. Potential antioxidant response to coffee—A matter of genotype? Meta Gene 2014, 2, 525–539. [Google Scholar] [CrossRef] [Green Version]
  146. Li, W.; Yue, W.; Zhang, L.; Zhao, X.; Ma, L.; Yang, X.; Zhang, C.; Wang, Y.; Gu, M. Polymorphisms in GSTM1, CYP1A1, CYP2E1, and CYP2D6 are Associated with Susceptibility and Chemotherapy Response in Non-small-cell Lung Cancer Patients. Lung 2012, 190, 91–98. [Google Scholar] [CrossRef]
  147. Sulem, P.; Gudbjartsson, D.; Geller, F.; Prokopenko, I.; Feenstra, B.; Aben, K.K.; Franke, B.; den Heijer, M.; Kovacs, P.; Stumvoll, M.; et al. Sequence variants at CYP1A1–CYP1A2 and AHR associate with coffee consumption. Hum. Mol. Genet. 2011, 20, 2071–2077. [Google Scholar] [CrossRef]
  148. Amin, N.; Byrne, E.; Johnson, J.; Chenevix-Trench, G.; Walter, S.; Nolte, I.M.; Vink, J.; Rawal, R.; Mangino, M.; Teumer, A.; et al. Genome-wide association analysis of coffee drinking suggests association with CYP1A1/CYP1A2 and NRCAM. Mol. Psychiatry 2012, 17, 1116–1129. [Google Scholar] [CrossRef]
  149. McMahon, G.; Taylor, A.E.; Smith, G.D.; Munafò, M.R. Phenotype Refinement Strengthens the Association of AHR and CYP1A1 Genotype with Caffeine Consumption. PLoS ONE 2014, 9, e103448. [Google Scholar] [CrossRef] [Green Version]
  150. Nault, R.; Doskey, C.M.; Fader, K.A.; Rockwell, C.; Zacharewski, T.; Zacharewski, T.R. Comparison of Hepatic NRF2 and Aryl Hydrocarbon Receptor Binding in 2,3,7,8-Tetrachlorodibenzo-p-dioxin–Treated Mice Demonstrates NRF2-Independent PKM2 Induction. Mol. Pharmacol. 2018, 94, 876–884. [Google Scholar] [CrossRef] [Green Version]
  151. Van den Bogaard, E.H.; Bergboer, J.G.; Vonk-Bergers, M.; Van Vlijmen-Willems, I.M.; Hato, S.V.; Van Der Valk, P.G.; Schröder, J.M.; Joosten, I.; Zeeuwen, P.L.; Schalkwijk, J. Coal tar induces AHR-dependent skin barrier repair in atopic dermatitis. J. Clin. Investig. 2013, 123, 917–927. [Google Scholar] [CrossRef] [Green Version]
  152. Zhao, Y.; Bao, R.-K.; Zhu, S.-Y.; Talukder, M.; Cui, J.-G.; Zhang, H.; Li, X.-N.; Li, J.-L. Lycopene prevents DEHP-induced hepatic oxidative stress damage by crosstalk between AHR–Nrf2 pathway. Environ. Pollut. 2021, 285, 117080. [Google Scholar] [CrossRef] [PubMed]
  153. Mukherjee, A.; Agrawal, S. A Global Perspective of Fine Particulate Matter Pollution and Its Health Effects. Rev. Environ. Contam. Toxicol. 2018, 244, 5–51. [Google Scholar] [CrossRef]
  154. Kwon, H.-S.; Ryu, M.H.; Carlsten, C. Ultrafine particles: Unique physicochemical properties relevant to health and disease. Exp. Mol. Med. 2020, 52, 318–328. [Google Scholar] [CrossRef] [PubMed]
  155. Dubowsky, S.D.; Suh, H.; Schwartz, J.; Coull, B.A.; Gold, D.R. Diabetes, Obesity, and Hypertension May Enhance Associations between AirPollution and Markers of Systemic Inflammation. Environ. Health Perspect. 2006, 114, 992–998. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Rückerl, R.; Hampel, R.; Breitner, S.; Cyrys, J.; Kraus, U.; Carter, J.; Dailey, L.; Devlin, R.B.; Diaz-Sanchez, D.; Koenig, W.; et al. Associations between ambient air pollution and blood markers of inflammation and coagulation/fibrinolysis in susceptible populations. Environ. Int. 2014, 70, 32–49. [Google Scholar] [CrossRef] [PubMed]
  157. Haberzettl, P.; Conklin, D.J.; Abplanalp, W.T.; Bhatnagar, A.; O’Toole, T.E. Inhalation of Fine Particulate Matter Impairs Endothelial Progenitor Cell Function Via Pulmonary Oxidative Stress. Arter. Thromb. Vasc. Biol. 2018, 38, 131–142. [Google Scholar] [CrossRef] [Green Version]
  158. Gao, M.; Ma, Y.; Luo, J.; Li, D.; Jiang, M.; Jiang, Q.; Pi, J.; Chen, R.; Chen, W.; Zhang, R.; et al. The Role of Nrf2 in the PM-Induced Vascular Injury Under Real Ambient Particulate Matter Exposure in C57/B6 Mice. Front. Pharmacol. 2021, 12, 618023. [Google Scholar] [CrossRef]
  159. Ge, C.; Hu, L.; Lou, D.; Li, Q.; Feng, J.; Wu, Y.; Tan, J.; Xu, M. Nrf2 deficiency aggravates PM2.5-induced cardiomyopathy by enhancing oxidative stress, fibrosis and inflammation via RIPK3-regulated mitochondrial disorder. Aging 2020, 12, 4836–4865. [Google Scholar] [CrossRef]
  160. Canova, C.; Dunster, C.; Kelly, F.J.; Minelli, C.; Shah, P.L.; Caneja, C.; Tumilty, M.K.; Burney, P. PM10-induced Hospital Admissions for Asthma and Chronic Obstructive Pulmonary Disease: The modifying effect of individual characteristics. Epidemiology 2012, 23, 607–615. [Google Scholar] [CrossRef]
  161. Wittkopp, S.; Staimer, N.; Tjoa, T.; Stinchcombe, T.; Daher, N.; Schauer, J.J.; Shafer, M.M.; Sioutas, C.; Gillen, D.L.; Delfino, R.J. Nrf2-related gene expression and exposure to traffic-related air pollution in elderly subjects with cardiovascular disease: An exploratory panel study. J. Expo. Sci. Environ. Epidemiol. 2016, 26, 141–149. [Google Scholar] [CrossRef] [Green Version]
  162. Hou, W.; Hu, S.; Li, C.; Ma, H.; Wang, Q.; Meng, G.; Guo, T.; Zhang, J. Cigarette Smoke Induced Lung Barrier Dysfunction, EMT, and Tissue Remodeling: A Possible Link between COPD and Lung Cancer. BioMed Res. Int. 2019, 2019, 2025636. [Google Scholar] [CrossRef] [PubMed]
  163. Giebe, S.; Hofmann, A.; Brux, M.; Lowe, F.; Breheny, D.; Morawietz, H.; Brunssen, C. Comparative study of the effects of cigarette smoke versus next generation tobacco and nicotine product extracts on endothelial function. Redox Biol. 2021, 47, 102150. [Google Scholar] [CrossRef] [PubMed]
  164. Yu, B.; Chen, J.; Liu, D.; Zhou, H.; Xiao, W.; Xia, X.; Huang, Z. Cigarette Smoking Is Associated with Human Semen Quality in Synergy with Functional NRF2 Polymorphisms. Biol. Reprod. 2013, 89, 5. [Google Scholar] [CrossRef]
  165. Masuko, H.; Sakamoto, T.; Kaneko, Y.; Iijima, H.; Naito, T.; Noguchi, E.; Hirota, T.; Tamari, M.; Hizawa, N. An interaction between Nrf2 polymorphisms and smoking status affects annual decline in FEV1: A longitudinal retrospective cohort study. BMC Med. Genet. 2011, 12, 97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Function and structure of Nrf2. (A) Under oxidative stress, Nrf2 dissociates from Keap1, to be translocated to the nucleus where heterodimerizes with the sMAF proteins for recognition of ARE within several genes to induce their expression. (B) Nrf2 consists of seven domains named Neh1 to 7, which allow specific interaction for its regulation. ARE, antioxidant response elements; β-TrCP, beta-transducin repeat-containing protein; Brg1, Brahma-related gene 1; CBP, CREB binding protein; CDH6, chromo-ATPase/helicase DNA binding protein 6; GPx, glutathione peroxidase; GST, glutathione-S-transferase (GST); HO-1, heme-oxygenase 1; Keap1, Kelch-like ECH-associated protein 1; NQO1, NAD(P)H quinone oxidoreductase 1; Neh, Nrf2-ECH homology; Nrf2, nuclear factor erythroid 2-related factor 2; Prdx, peroxiredoxins; RXRα, retinoid X receptor alpha; SOD, superoxide dismutase; sMAF, small musculoaponeurotic fibrosarcoma. Figure created with BioRender at biorender.com (accessed on 10 February 2022).
Figure 1. Function and structure of Nrf2. (A) Under oxidative stress, Nrf2 dissociates from Keap1, to be translocated to the nucleus where heterodimerizes with the sMAF proteins for recognition of ARE within several genes to induce their expression. (B) Nrf2 consists of seven domains named Neh1 to 7, which allow specific interaction for its regulation. ARE, antioxidant response elements; β-TrCP, beta-transducin repeat-containing protein; Brg1, Brahma-related gene 1; CBP, CREB binding protein; CDH6, chromo-ATPase/helicase DNA binding protein 6; GPx, glutathione peroxidase; GST, glutathione-S-transferase (GST); HO-1, heme-oxygenase 1; Keap1, Kelch-like ECH-associated protein 1; NQO1, NAD(P)H quinone oxidoreductase 1; Neh, Nrf2-ECH homology; Nrf2, nuclear factor erythroid 2-related factor 2; Prdx, peroxiredoxins; RXRα, retinoid X receptor alpha; SOD, superoxide dismutase; sMAF, small musculoaponeurotic fibrosarcoma. Figure created with BioRender at biorender.com (accessed on 10 February 2022).
Antioxidants 11 00507 g001
Figure 2. Regulatory mechanisms of Nrf2. (A) For Nrf2 regulation at protein level, Nrf2 is ubiquitinated (Ub) for its proteasomal degradation through Keap1-dependent and β-TrCP-dependent mechanisms. (B) For Nrf2 activation, the canonical pathway involves the chemical modification of Keap1 cysteines (Cys) that elicits the release of Nrf2 for its nuclear translocation. Additionally, phosphorylation (P) by cytosolic kinases promotes Nrf2 activation. AMPK, AMP-activated protein kinase; β-TrCP, beta-transducin repeat-containing protein; CK2, casein kinase 2; Cul3, cullin 3; GSK-3, glycogen synthase kinase 3; Keap1, kelch-like ECH-associated protein 1; Nrf2, nuclear factor erythroid 2-related factor 2; PKC, protein kinase C; Rbx, ring box-1; ROS, reactive oxygen species; SFC, Skp1–Cullin–F-box protein complex. Figure created with BioRender at biorender.com (accessed on 10 February 2022).
Figure 2. Regulatory mechanisms of Nrf2. (A) For Nrf2 regulation at protein level, Nrf2 is ubiquitinated (Ub) for its proteasomal degradation through Keap1-dependent and β-TrCP-dependent mechanisms. (B) For Nrf2 activation, the canonical pathway involves the chemical modification of Keap1 cysteines (Cys) that elicits the release of Nrf2 for its nuclear translocation. Additionally, phosphorylation (P) by cytosolic kinases promotes Nrf2 activation. AMPK, AMP-activated protein kinase; β-TrCP, beta-transducin repeat-containing protein; CK2, casein kinase 2; Cul3, cullin 3; GSK-3, glycogen synthase kinase 3; Keap1, kelch-like ECH-associated protein 1; Nrf2, nuclear factor erythroid 2-related factor 2; PKC, protein kinase C; Rbx, ring box-1; ROS, reactive oxygen species; SFC, Skp1–Cullin–F-box protein complex. Figure created with BioRender at biorender.com (accessed on 10 February 2022).
Antioxidants 11 00507 g002
Table 1. Association of SNPs of Nrf2 and related genes with cardiometabolic diseases.
Table 1. Association of SNPs of Nrf2 and related genes with cardiometabolic diseases.
GeneSNP IDStudy GroupFindingReference
Nrf2rs2364723SAPALDIA cohort
2002/2003 (N = 1472)
2010/2011 (N = 1235)
Adverse obesity effects on heart rate variability.[80]
Nrf2rs6721961Mexican patients: N = 627 diabetic subjects and N = 1020 controls; Chinese patients: N883 T2DM subjects (383 with diabetic nephropathy).Associated with T2DM development and complications. Higher risk for CAD and severity of coronary atherosclerosis.[94,95,108]
Nrf2rs35652124Indian subjects N = 400 (100 with diabetic foot ulcer, 150 T2DM patients, and 150 healthy subjects).
N = 464 Japanese subjects (285 men and 179 women).
N64 healthy African American.
N = 184 white individuals
Association with insulin resistance and diabetic complications.
Cardiovascular disease-associated mortality. Impairment of vasodilation.
[97,104,122]
Nrf2rs2364723SAPALDIA cohort 2002/2003 (N = 1472)
2010/2011 (N = 1235).
N = 1390 subjects from the Vlagtwedde–Vlaardingen cohort.
N = 809 Chinese volunteers (214 T2DM patients, 236 without diabetic complications, and 359 healthy individuals).
Impairment of cardiovascular function. Low risk of mortality in cardiovascular diseases. Diabetic complications.
Associated with lower triglyceride levels.
[59,80,89]
Nrf2rs13001694N = 1390 subjects from the Vlagtwedde–Vlaardingen cohort.Associated with low risk of mortality in cardiovascular diseases.[59]
Nrf2rs182428269N = 400 (150 2DM subjects, 150 healthy subjects, and 100 with diabetic complications).Risk for T2DM development and diabetic complications[98]
Nrf2rs10497511N = 809 Chinese volunteers (214 T2DM patients, 236 without complications, and 359 healthy individuals).Diabetic complications.[89]
Nrf2rs1962142N = 809 Chinese volunteers (214 T2DM patients, 236 without diabetic complications, and 359 healthy individuals).Diabetic complications.[89]
Nrf2rs6726395N = 809 Chinese volunteers (214 T2DM patients, 236 without diabetic complications, and 359 healthy individuals).Diabetic complications.[89]
KEAP1rs11085735N = 117 patients with fatal and nonfatal cardiovascular events (42 died).Predictor of cardiovascular events.[57]
HMOX1rs2364723N = 809 Chinese volunteers (214 T2DM patients, 236 without diabetic complications, and 359 healthy individuals).Negative associated with diabetic complications.[89]
GPxrs1050450416 Mexican women (N = 208 healthy subjects, N = 208 obese subjects).
N = 1244 T2DM subjects and N = 730 healthy subjects.
N = 773 Caucasian subjects genotyped from the UCL Diabetes and Cardiovascular disease Study and N = 382 Caucasian subjects from the Ealing Diabetes Study.
N = 396 T2DM patients and N = 678 control subjects.
Morbid obesity development, particularly in females. Associated with diabetic neuropathy and development of carotid plaques in patients with diabetes.[90,91,92,93]
NQO1rs1800566N = 2374 Thai subjects (with and without CAD).
N = 834 T2DM patients (601 Seoul set and 233 Koyang set).
N = 130 patients (67 patients with coronary heart disease and 63 healthy individuals).
Associated with severity of coronary atherosclerosis and CAD in females. Controversial association with cardiovascular risk. Associated with hypertension.[108,116,117]
CATrs1049982N = 1388 participants > 18 years old (704 women, 300 untreated hypertensive patients).Risk of hypertension.[125]
TXNrs2301241N = 1388 participants > 18 years old (704 women, 300 untreated hypertensive patients).Associated with high blood pressure.[125]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zazueta, C.; Jimenez-Uribe, A.P.; Pedraza-Chaverri, J.; Buelna-Chontal, M. Genetic Variations on Redox Control in Cardiometabolic Diseases: The Role of Nrf2. Antioxidants 2022, 11, 507. https://0-doi-org.brum.beds.ac.uk/10.3390/antiox11030507

AMA Style

Zazueta C, Jimenez-Uribe AP, Pedraza-Chaverri J, Buelna-Chontal M. Genetic Variations on Redox Control in Cardiometabolic Diseases: The Role of Nrf2. Antioxidants. 2022; 11(3):507. https://0-doi-org.brum.beds.ac.uk/10.3390/antiox11030507

Chicago/Turabian Style

Zazueta, Cecilia, Alexis Paulina Jimenez-Uribe, José Pedraza-Chaverri, and Mabel Buelna-Chontal. 2022. "Genetic Variations on Redox Control in Cardiometabolic Diseases: The Role of Nrf2" Antioxidants 11, no. 3: 507. https://0-doi-org.brum.beds.ac.uk/10.3390/antiox11030507

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