Skip to main content

The impact of reactive oxygen species in the development of cardiometabolic disorders: a review

Abstract

Oxidative stress, an alteration in the balance between reactive oxygen species (ROS) generation and antioxidant buffering capacity, has been implicated in the pathogenesis of cardiometabolic disorders (CMD). At physiological levels, ROS functions as signalling mediators, regulates various physiological functions such as the growth, proliferation, and migration endothelial cells (EC) and smooth muscle cells (SMC); formation and development of new blood vessels; EC and SMC regulated death; vascular tone; host defence; and genomic stability. However, at excessive levels, it causes a deviation in the redox state, mediates the development of CMD. Multiple mechanisms account for the rise in the production of free radicals in the heart. These include mitochondrial dysfunction and uncoupling, increased fatty acid oxidation, exaggerated activity of nicotinamide adenine dinucleotide phosphate oxidase (NOX), reduced antioxidant capacity, and cardiac metabolic memory. The purpose of this study is to discuss the link between oxidative stress and the aetiopathogenesis of CMD and highlight associated mechanisms. Oxidative stress plays a vital role in the development of obesity and dyslipidaemia, insulin resistance and diabetes, hypertension via various mechanisms associated with ROS-led inflammatory response and endothelial dysfunction.

Introduction

Cardiometabolic disorders (CMD) is a constellation of metabolic predisposing factors for atherosclerosis such as insulin resistance (IR) or diabetes mellitus (DM), systemic hypertension, central obesity, and dyslipidaemia [1]. They contribute to the global death rate and remain a public health challenge. There is a significant rise in the prevalence of CMD not only in high-resource countries but also in developing nations with emerging economies [2, 3]. Although, there are available data on the development of CMD and the mechanisms associated with its attendant complications, novel mechanisms are still revealed by recent studies in an attempt to open new therapeutic opportunities [4, 5]. Several studies have implicated oxidative stress in CMD development. It has been reported that cardio-tolerance to oxidative stress reduces with advancing age due to the antioxidant levels, particularly enzymatic antioxidants, contributing to the development of CMD [6]. Also, this has been linked with arterial thickening and atherosclerosis [7, 8]. This ensues in vascular endothelial damage and remodelling.

The high prevalence of CMD is a global phenomenon. The increase in the global prevalence is seemingly due to a parallel rise in the incidence of dietary and lifestyle changes, and cases of obesity [9]. There is an anticipated increase in these disorders due to projections of a greater incidence in future obesity cases [9]. The increased incidence of CMD among urban women when compared with their rural counterparts has been attributed to the global increase in urbanization and decline in physical exercise, especially in Africa [10]. A Nigerian study reported a prevalence of 18.0% and 10.0% in the semi-urban and rural community respectively and 34.7% and 24.7% respectively in a hypertensive population [11]. In Tunisia, Hosseinpanah et al. [12] documented a prevalence of 55.8% and 30.0% in women and men respectively. The higher prevalence in women was ascribed to the lower high-density lipoprotein (HDL), higher incidence of central obesity, and hypertension. According to the findings of Harzallah in Turkey [13], though female had a higher prevalence (39.6%) than the male counterpart (28%), the prevalence was similar in the urban (33.8%) and rural (33.9%) settings. In Qatar, prevalence rate of 26.5% was reported using Adult Treatment Panel III (ATP III) criteria and 33.7% using International Diabetes Foundation (IDF) criteria [14]. The observed incidence rose with advancing age and increasing body mass index, but reduced advancement in education and regular physical activity. In Lebanon, a prevalence of 31.2% was reported with men having a striking higher tendency [15]. Sibai [16] documented an age-adjusted prevalence of 37% in males in Saudi Arabia, with a higher prevalence in male (44%) than female (35.6%) [16]. In the USA in 2003/2004, using the National Cholesterol Education Program (NCEP)/ATP III criteria, about 34% people above 20 years old had CMD [17], with a marginal higher prevalence in male (35.1%) than female (32.6%). Although the prevalence of the disorder varies across geographic regions and age groups, about 25% of the adult European population was reported to have CMD [18].

This review highlights the role of ROS in the pathogenesis of CMD and discusses the associated mechanisms. This will shed more light to the pathogenesis of CMD and consequent open new therapeutic horizons.

Methods

The present study reviewed all available data published in peer-reviewed journals up till date. Search was made using AJOL, DOAJ, Embase, Google Scholar, Pubmed/Pubmed Central, and Scopus databases using relevant key word searches like “Cardiometabolic disorders”, “metabolic disorders”, “Reactive oxygen species”, “ROS”, “oxidative stress”, “lipid peroxidation”, “Nitrosative stress”, and “Antioxidants and cardiometabolic disorders”. Papers published in peer-reviewed journals were included in this narrative review. Papers that did not adequately discuss details of the study were excluded from this review. Duplicated records were also excluded.

Discussion

Pathophysiology of CMD

CMD involve interplay of a cascade of pathophysiological events ensuing in a rise in IR, accumulation of free fatty acids (FFA) in the circulation, lipid and glucose dysmetabolism, and raised levels of adipokines and cytokines [19,20,21]. Since insulin controls adipose tissue lipid breakdown, the primary source of plasma (FFA), excess visceral fat causes IR with a resultant increase in lipids breakdown [22]. IR is further triggered by the increasing FFA concentrations via enhanced glucose dysregulation [23, 24]. These cumulate in ladening of fatty deposits in the blood vessels with resultant vasoconstriction, excessive fluid retention, and sustained rise in blood pressure [22].

A raised level of FFA does not just prevent the stimulation of glucose uptake in the muscle by insulin, [25] it also depresses the production of glucose in the liver [26], and enhances hepatic uptake of FFA. It causes increased production of VLDL and triglyceride (TG) in the liver [27,28,29,30], thus promoting TG transfer from VLDL to HDL and subsequent clearance of HDL [29, 30].

Inflammatory cytokines are generated by the adipose tissue and have the potential to trigger IR and adiponectin [31, 32]. Tumour necrosis factor-alpha (TNF-α) suppresses insulin signalling [33], interleukin-6 (IL-6) directly induces inflammation or enhances the release of hepatic C-reactive protein [34], while interleukin-8 (IL-8) activates neutrophil granulocytes. Adiponectin increases hepatic insulin sensitivity and oxidation of skeletal muscle glucose and fatty acid, and decreases glucose release [35,36,37,38]. This may infer that when there is abnormal plasma FFA concentration, the production of adipokines is raised while that of adiponectin is reduced.

Though obesity-associated excess visceral and/or intraperitoneal fat is strongly linked with IR [25, 26, 39, 40], whether or not intraperitoneal fat causes or is just associated with IR remains unclear. However, studies have suggested that fatty acids from lipolysis of intraperitoneal fat are a vital influential predisposing factor of IR since they are delivered to the liver directly through the portal vein [41]. Findings from the study of Havel et al. [42] among obese subjects revealed that lipolysis of the intraperitoneal fat accounts for 20% of FFA delivered to the liver and 15% delivered to skeletal muscle. Thus, intraperitoneal fat possibly contributes to hepatic IR, but unlikely to trigger skeletal muscle IR.

The pathogenesis of CMD involves a complex interplay between genetics and environmental factors. Also, epigenetic factors such as DNA methylation and histone modification are possibly key in the incidence of these disorders by mediating the effects of environmental exposures on the risk of development of CMD. The genetic factors are primarily hereditary and non-modifiable, while the environmental factors are modifiable. Race and advancing age are additional non-modifiable factors. Table 1 shows a list of some modifiable factors. Individuals at risk could be influenced by one or more hereditary and environmental factors which may worsen the pathogenic progress [22]. Thus, lifestyle modifications like diet and weight control, optimal exercise, cessation of cigarette smoking, and control of pollution and exposure to other mitochondrial toxins are beneficial factors that assuage CMD development.

Table 1 Modifiable protective and risk factors for cardiometabolic disorders

Epigenetics are alterations caused by developmental processes or environmental influences that do not modify the genetic code but influences the expression of the information encoded in the DNA [43]. Though there is limited evidence from genetic studies for a common genetic soil for CMD development, and contradicting relationships of genes and gene variants exists [44]. In order to promote early detection, prompt management and likely preventive strategies, a good an in-depth knowledge of the predisposing genetic factors influencing the development of CMD is important.

Oxidative stress

Oxidative stress (OS) is the presence of reactive oxygen species (ROS) in excess of the antioxidant buffering capacity. It can be described as an imbalance in ROS generation and antioxidant defence leading to accumulation of ROS. It is an alteration in the balance of prooxidant /antioxidant system in favour of prooxidant with attendant lethal effects with resultant damage to the cellular macromolecules (Fig. 1). In organisms, including humans, ROS and free radicals are produced during metabolic and immune system function. Molecular oxygen (O2) can unpair and leave free radicals which are highly unstable and reactive, leading to the formation of ROS [45, 46].

Fig. 1
figure 1

Oxidative stress resulting from an imbalance between ROS generation and antioxidant system and its consequences on cellular macromolecules

Free radicals are molecules with at least one unpaired electrons in their outer orbit. This makes them highly unstable and reactive with other molecules to produce more stable species [47]. Such radicals include Reactive Oxygen Species (ROS) and Reactive Nitrogen Species (RNS) [48]. Free radicals and oxidant species are involved in cellular and tissue dysfunction, and thus toxic [49]. Although low or moderate concentrations of these molecules play physiological roles such as signalling processes and defence mechanisms against infectious agents [49], when they are generated in excess, they lead to lipid, protein and DNA damage. This is a known fact, though stunning that ROS protects the cell against ROS damage by stimulating antioxidant responses and maintaining or re-establishing redox balance. It is noteworthy to state that in non-pathological states, ROS in low to moderate concentrations play a key homeostatic function in cellular and mitochondrial signaling and functionality; but, in excess concentrations when unchecked, it mediates oxidative cell and tissue damage. This can trigger positive feedback [50].

Sources of cardiovascular ROS

Sources of ROS in cells

Vascular cells, including cardiac cells and neurons, generate ROS, thus triggering the incidence of CMD. Various enzyme systems such as cytochrome P450, the mitochondrial respiratory chain, xanthine oxidase (XO), uncoupled endothelial nitric oxide synthase (eNOS), heme oxygenase (HO), myeloperoxidase (MPO), lipoxygenase (LOX), cyclooxygenase (COX) and NADPH oxidases (NOX) generate ROS especially in a pathological state [51].

XO is a xanthine oxidoreductase that exists in two forms; xanthine dehydrogenase (XDH), which is the predominant form and can be irreversibly or reversibly converted into XO via proteolysis or the oxidation of cysteine residues respectively [52]. Its expression in the vascular endothelium is promoted by angiotensin II (Ang II) or oscillatory shear in a NOX-dependent manner [53].

The NOX family includes NOX1-5 and DUOX1-2. In humans, NOX2 NADPH oxidase seems to be the most important source of ROS generation [54,55,56]. NADPH oxidases are specific source of ROS because they produce ROS in a tightly-controlled manner unlike as in other sources where ROS are produced as a secondary metabolite [57]. Also, NADPH oxidases can also generate ROS from other enzyme systems [57].

Superoxide radical (O2-) is the first moiety that is produced by most of the enzyme systems, particularly NADPH oxidases. O2- radical can undergo rapid dismutation to hydrogen peroxide (H2O2), which is mediates most signaling effects of ROS [57].

ROS in the heart

Excessive ROS generation in the mitochondria has been shown in cardiomyocytes [58]. Mitochondria are the powerhouses of the living cells and produce energy primarily via oxidative phosphorylation. Also, mitochondria are the major source of ROS in the cardiovascular system [51, 58] (Fig. 2, Table 2). The aconitase of the Kreb’s cycle in the mitochondria matrix produces NADH and FADH2 which are oxidized for ATP production in the electron transport chain (ETC) located in the inner membrane of the mitochondria. The ETC mediates electron flow through series of electron carriers including complexes I, II, III and IV as well as ubiquinone and cytochrome c. As electron flows through the ETC, protons are translocated from the mitochondria matrix to the mitochondria inter-membrane space, thus creating an electrochemical potential gradient across the inner membrane. Generation of superoxide anion radicals in the mitochondria is mainly by electron leakage from the chain. Under physiological conditions, the oxygen tension in the mitochondria is low in state 4 respiration and oxygen consumption by the chain does not meet the demand of oxidative phosphorylation. Hence, a decrease in the rate of mitochondrial oxidative phosphorylation increases electron leakage from the ETC and conversely generation of superoxide anion radical.

Fig. 2
figure 2

Generation of reactive oxygen species (ROS) by the mitochondria electron transport chain. Δp : proton motive force, ΔΨ: membrane potential, ΔpH : proton gradient

Table 2 Enzymes and inhibitors of the electron transport chain (ETC)

XO generates O2- as a secondary metabolite of purine catabolism. NOS uncoupling and subsequent O2- generation have been linked with vascular endothelial dysfunction. Although, NOS primarily produces nitric oxide (NO), it may generate O2- if it becomes uncoupled. Uncoupling of NOS is commonly seen when its co-factor, tetrahydrobiopterin (BH4), or its substrate, L-arginine, is deficient [59]. NADPH oxidases may also contribute to the generation of ROS by degrading BH4 via oxidation, thus causing NOS uncoupling [60], or activating xanthine oxidase [61].

Oxidative stress triggers apoptosis via several pathways activating enzymes involved in pro-apoptotic signalings, such as JNK, p38, ASK-1, and CaMKII [62] with resultant release of cytochrome-c. Though the excessive generation of ROS by NOX is deleterious, at minimal to moderate levels, H2O2 and O2- produced by NOX, act as signalling molecules, thus mediating physiological responses [63].

ROS and the vasculature

ROS influence pathological and physiological processes in the vasculature. Predisposing factors to CMD like DM, obesity, hypertension, dyslipidaemia, and ageing result in vascular dysfunction partly through oxidative stress [64]. ROS is essential for the growth, proliferation, and migration of endothelial cells (ECs) and smooth muscle cells (SMCs), as well as angiogenesis, apoptosis of EC and SMC, vascular tone, host defence, and genomic stability [65]. OS does not only induce macromolecular damage, it also alters vascular redox-dependent signaling pathways [66]. ROS target signaling pathways such as mitogen activated protein kinases (MAPKs) which include extracellular signal-regulated kinases (ERK1/2), p38 and c-Jun N-terminal kinases [67]. It also targets serine/threonine kinase Akt/protein kinase B, epidermal growth factors (EGF), and platelet-derived growth factors (PDGF). In oxidative stress, ROS triggers endothelial dysfunction via disruption of vasoprotective NO signaling [68].

NO is a vasodilator that acts via cGMP which is produced by SMCs. NO inhibits platelet adhesion and aggregation, thus acting as an anti-atherogenic factor. It also prevents leukocyte-endothelial interactions and MC proliferation [69]. O2- rapidly reacts with NO to produce peroxynitrite (ONOO-), which is also a potent oxidant [70]. ONOO- formation occurs more rapidly than superoxide dismutase (SOD)-dependent dismutation of O2- [71]. The formed ONOO- induces BH4 oxidation with resultant eNOS uncoupling [72] thus converting eNOS into a pro-oxidant [73]. Hence, ONOO- formation depletes NO concentration and also induces eNOS uncoupling.

ROS also triggers the structure of the inflammasome, IL-1β, and IL-8 via the activation of caspase-1 [74]. This pathway is vital in atherosclerosis.

As highlighted previously, NADPH oxidase, XO, and uncoupled eNOS, including lipoxygenase, cyclooxygenase, and cytochrome P450 monooxygenase are key enzymes that generate ROS in the cells as well as in the vascular wall. ROS are involved in the structural modification of the vascular wall thickness and lumen diameter [75]. This is consequent to the passive adaptation to chronic changes in hemodynamics and neuro-humoral factors such as angiotensin II and ROS [75]. This vascular remodelling could be inward eutrophic or hypertrophic, and play a role in the development of hypertension. The inward eutrophic remodelling involves reduction in the size of the lumen, media thickening, improved media to lumen ratio, and little alteration in the cross-sectional area of the media [76], while the hypertrophic remodelling involves an enhancement in the cross-sectional area of the vascular wall, size of the cell and accumulation of ECM proteins like collagen and fibronectin [77].

Vascular ROS scavenging

Mitochondria as the primary source of ROS

CMD is accompanied by an imbalance in vascular ROS generation and scavenging. The primary defense against vascular ROS is discussed below.

Superoxide dismutase

In humans, SOD1, SOD2, and SOD3 are the three known SOD isoforms. SOD1 (copper, zinc [Cu-Zn]-SOD) is located in the cytoplasm and mitochondrial intermembrane space, SOD2 (Mn-SOD) is located in the mitochondrial matrix, while SOD3 (extracellular [EC]-SOD) is located in the extracellular space [78]. SOD dismutates O2- to H2O2 and oxygen, hence prevents the inactivation of NO. However, high concentrations of the secondary metabolites of the dismutation may cross cellular membranes to generate pro-atherogenic molecules [78,79,80,81]. The effect of SOD is dose-dependent.

Catalase

Catalase decomposes H2O2 to oxygen and water. Up-regulation of this enzymatic antioxidant inhibits atherosclerosis [80] and impairs angiotensin II-mediated aortic wall hypertrophy [82].

Glutathione peroxidase

Glutathione peroxidase (GPx) catalyzes the reduction of H2O2 to water, and lipid peroxides to their alcohols [83, 84]. Although there are 8 isoforms of GPx, GPx1 and GPx4 seem to be the most studied. GPx1 is found in many cell types, and its deficiency has been linked with atherosclerosis [85, 86]. On the other hand, GPx4 is expressed in the endoplasmic reticulum (ER), cytoplasm, mitochondria, and plasma membrane. GPx4 prevents atherogenesis by impairing lipid peroxidation and the sensitivity of vascular cell to oxidized lipids [87].

Paraoxonase

Paraoxonases (PON) include PON-1, PON-2, and PON-3. PON exhibit anti-atherogenic properties, possibly via inhibition of oxidative stress [88]. PON-1 is primarily secreted by the liver [89]. PON-1 prevents the peroxidation of HDL and LDL, breaks down cholesteryl esters and lipoproteins seen in oxidized lipoproteins, and also prevents OS, inflammation, and monocyte attraction via blunting myeloperoxidase-induced ROS generation. PON-2, which is expressed in the ER and mitochondrial membranes, exerts its effects on the vascular cells [90, 91]. PON-3, which is located in the serum and cells [92], prevents atherogenesis [93, 94].

Heme oxygenase

HO catalyzes degradation of heme to carbon monoxide (CO), biliverdin, and free ferrous iron [95]. HO exists in three isoforms; the inducible (HO-1), constitutive (HO-2), an enzymatically inactive (HO-3) forms [95]. OS, hypoxia, and some cytokines stimulate the upregulation of the inducible isoform, which are key in impairing vascular remodelling and atherosclerosis [95]. Although at moderate concentrations, the CO produced by HO has anti-inflammatory, antiproliferative, and vasodilatory activities, it is toxic at a very high concentration [96]. Biliverdin is a pigment that scavenges radicals and also blunts the effect of NOX [97].

Thioredoxin

Thioredoxin (Trx) is an enzymatic antioxidant that is located in the ECs, SMCs, and fibroblasts [98]. It is vasoprotective and reverses age-related arterial stiffness and raised blood pressure via enhancement of vascular redox and restoration of the function of eNOS [98].

Non-enzymatic antioxidants

Bilirubin, uric acid, glutathione, exogenous substances like vitamins (mainly vitamins C and E) and polyphenols, contribute to the antioxidant defence system [99]. Bilirubin and uric acid scavenge extracellular radicals, while glutathione modifies the intracellular redox state [99]. Vitamin C (ascorbic acid) scavenges several oxidative/nitrogen species, stabilizes BH4 and eNOs, and restores vitamin E from its radical state (tocopheroxyl radical) [100]. α-tocopherol is the principal member of vitamin E with antioxidant property. Sources of polyphenolic antioxidants include food such as vegetables and cocoa and beverages. Polyphenolics impair NADPH oxidases activities [101, 102].

Micronutrients like selenium, copper, zinc, iron, and calcium also contribute to the antioxidant buffering capacity. Selenium protects against oxidative DNA damage. It acts via selenoenzymes-mediated mechanism such as GPx [103]. As selenium serves as a co-factor of GPx, copper and zinc are co-factors of SOD, while the iron is a co-factor of catalase. Thus, these elements influence the activities of the respective enzymatic antioxidants. Note worthily, iron and copper may act as pro-oxidants by catalyzing the production of hydroxyl (OH) radicals from O2- and H2O2 [104, 105]. Although calcium is critical in excitation-contraction coupling, it also plays vital physiological roles like the regulation of gene expression and cellular energetics [106,107,108,109,110].

Biomarkers of oxidative stress

Oxidative stress has been shown to be a key player in various diseases including cardiometabolic disorders. A wide range of methods have been developed and employed to measure the nature and extent of oxidative stress ranging from oxidation of lipids to free amino acids and proteins, and DNA. Although diverse oxidative stress biomarkers are available as predictors of various diseases, the specificity of each seems to be yet established. Available biomarkers of oxidative stress have been summarized in Table 3.

Table 3 Biomarkers of oxidative stress

Oxidative stress and CMD

Oxidative stress and obesity

Emerging evidences implicating OS as the soil for the initiation and progression of dyslipidaemia, obesity, IR, DM, hypertension and atherosclerosis exit. Obesity is the primary causal component of CMD [111,112,113,114] (Fig. 3). Consumption of an energy-dense meal is linked to a significant increase in the concentrations of 4-hydroxyl 2-nonenal (HNE) [116], thus is essential in the incidence of obesity. Notably, Johnson et al. [117] have reported a reduced HNE level of HNE in obese individuals on calorie restriction. This underscores the essence of OS via HNE in incident obesity. ROS generation rises in parallel with fat adipocyte fat accumulation, and the increase in the level of FFAs also stimulates adipocyte ROS generation via NADPH oxidase activation and decline in enzymatic antioxidant expression [115]. In the presence of OS in adipocytes, anti-inflammatory adiponectin level falls [118, 119], while pro-inflammatory adipocytokines concentration rises [115, 120, 121]. Dysregulation of adipocytokines is vital in the development of obesity-associated metabolic disorder. Raised adipocyte generation of PAI-1, MCP-1, and TNF-α is important in the pathogenesis of thrombosis [122], and IR [123, 124]. Since adiponectin increases cellular sensitivity to insulin [37, 125,126,127] and also possesses anti-atherogenic effects [128,129,130], a marked reduction in the circulatory concentrations of adiponectin results in IR and atherosclerosis via systemic inflammation [115]. HNE up-regulates the expression of inducible cyclooxygenase (COX-2) and PAI-1 [131] and down-regulates the expression of adiponectin [119, 132]. Dysregulation of adipocytes results in systemic inflammation; this as well as increased adipocyte ROS generation promotes endothelial dysfunction. This is key in the development of IR, DM and atherosclerosis. Interestingly, renin-angiotensin-aldosterone system (RAAS) which plays an important role in blood pressure and volume regulation, has also been demonstrated to trigger adipocyte ROS generation [133].

Fig. 3
figure 3

The role of oxidative stress in the development of obesity and cardiometabolic disorders. This illustration is a modification of the working model illustrating how increased ROS production in accumulated fat contributes to metabolic syndrome by Furukawa et al. [115]

Oxidative stress and IR/diabetes

Following systemic inflammation, oxidative damage to the endothelial cells causes impaired glucose uptake and utilization by hepatocytes and skeletal myocytes. Activation of NOX via RAAS increases endothelial ROS generation [134, 135]. This is dependent on angiotensin II type-1 and mineralocorticoid receptors [136]. This cascade of events continues and causes a transition from IR to DM type II [137]. Oxidative injury to the endothelium reduces the circulatory level of NO due to the decline in its synthesis by uncoupling eNOS via ROS-induced oxidation and depletion of BH4 [138, 139]. Increased generation of ONOO- via coupling of NO to superoxide also contributes to NO depletion [140]. The generated ONOO- is very reactive and leads to endothelial cell death [140, 141] which also impairs endothelial NO generation. It is a known fact that eNOS-derived NO is essential in angiogenesis by enhancing vascular endothelial growth factors and up-regulating the recruitment of endothelial progenitor cells from the bone marrow [142, 143]. Hence ROS-induced decline in circulatory NO secondary to endothelial dysfunction impairs the growth of the capillary network and blood flow regulation and subsequent diminution of microcirculation in metabolically active tissues and dysregulations of glucose and dyslipidaemia.

Studies have implicated OS in IR and DM through insulin signaling, insulin-induced GLUT 4 translocation and glucose uptake via insulin receptor substrate (IRS) phosphorylation, MARK activation and ER stress. Activation of serine/threonine kinase cascade stimulates serine phosphorylation of IRS that in turn impairs tyrosine phosphorylation thus possibly enhancing IRS degradation [144], resulting to alteration of glucose uptake signalling pathways by GLUT4 via IRS-1 and phosphatidylinositol 3-kinase (PI3K)/Akt [137]. Impairment of glucose uptake through this pathway exerts a further negative effect on IR. Adipocytes, which act as glucose sensors [145], senses impaired GLUT4-mediated glucose uptake and release adipocytokines (such as retinol-binding protein 4, RBP 4) to prevent glucose uptake by skeletal muscle and improve hepatic glucose output through insulin signalling blockade [146]. This sums up to raised plasma level of glucose. Hence OS mediates the development of IR and DM type II via downregulation of circulatory NO bioavailability and GLUT 4 expression in adipocytes (Fig. 4).

Fig. 4
figure 4

The role of oxidative stress in the pathogenesis of insulin resistance/type II diabetes. This illustration is a modification of the mechanism illustrating the role of oxidative stress to adipocytes in insulin resistance by Otani H [137].

Oxidative stress and hypertension

The role of NO in the pathophysiology of hypertension has well clearly described. NO enhances angiogenesis and blood supply, thus regulates blood pressure. NO may react with O- to increase the generation of ONOO-, which break down to form hydroxyl radical [147, 148]. The generated ONOO- causes eNOS uncoupling, iNOS uncoupling, and BH4 depletion [137]. Under physiological conditions, electrons are transferred from a heme group in the oxygenase domain to L-arginine by eNOS. This leads to the generation of L-citrulline and NO [149]. However, when there is a depletion of NO, may be due to depletion of L-arginine, a substrate of NO, or BH4, a co-factor in NO production, eNOS switches to an uncoupled state from a coupled state. This leads to the reduction of oxygen by electrons from the heme group with subsequent production of superoxide radical (O2-) [150]. This radical reacts with NO to further deplete NO, leading to endothelial dysfunction and rise peripheral resistance, thus causing a sustained rise in blood pressure. Depletion of L-arginine is associated with exaggerated arginase II activity. iNOS expression is also up-regulated in endothelial dysfunction [137]. Although iNOS generates NO, due to oxidative stress-induced BH4 depletion and iNOS uncoupling, iNOS uncoupling enhances oxidative stress. Also, it leads to a vicious cycle of endothelial dysfunction and persistent rise in blood pressure (Fig. 5). In addition, endogenous eNOS inhibitor, asymmetric dimethyl-L-arginine (ADMA) may contribute to eNOS uncoupling. Oxidative stress has been demonstrated to enhance the activity of protein arginine N-methyltransferase (PRMT) and reduce the activity of dimethylarginine dimethylaminohydrolase (DDAH), an ADMA-degrading enzyme, thus resulting in a rise in ADMA concentrations [69, 151, 152]. The rise in ADMA levels may inhibit NO synthesis by eNOS or even lead to eNOS uncoupling [69, 151]. Note worthily, BH4 depletion-driven eNOS uncoupling may rather be secondary to the oxidation of zinc-thiolate cluster of eNOS. Exposure of isolated eNOS to ONOO- causes disruption of the zinc-thiolate cluster of the enzyme [153]. Since the Cys99 in this cluster is essential for BH4 binding, oxidation of this cluster will disrupt the BH4 binding site of the enzyme; an state similar to depletion of BH4.

Fig. 5
figure 5

The role of oxidative stress in the pathogenesis of hypertension. XO: xanthine oxidase, NOX: nicotinamide adenine dinucleotide phosphate oxidase, eNOS: endothelial nitric oxide synthase, OxLDL: Oxidized LDL, ROS: reactive oxygen species, PDGF: Platelet-derived growth factor, SMC: smooth muscle cells, MF: Myofibrils

Local RAAS activation plays an integral role in the complex cascades that contribute to endothelial dysfunction. Studies have shown that the accumulation of visceral fat and raised OS and inflammatory response in adipose tissue enhance the release of components of adipose RAAS. Animal studies have demonstrated up-regulation of angiotensinogen in fatty tissue in obesity. This is strongly linked with hypertension [154]. Angiotensinogen converts angiotensin I to angiotensin II, which exerts its effects via angiotensin II type 1 receptor (AT1R) [149]. In the zona glomerulosa of the adrenal cortex, activation of AT1R triggers the release of mineralocorticoids [149], which elevate blood pressure primarily via stimulation of sodium reabsorption and expansion of plasma volume, and secondarily via non-genomic mineralocorticoid receptor (MR)-mediated actions [155]. In non-adrenal tissues, activation of AT1R triggers ROS generation with resultant impairment of insulin signaling, as well as proliferative and inflammatory responses [156]. This penultimately leads to endothelial dysfunction and hypertension. It is worthy to note that mineralocorticoids such as aldosterone and deoxycorticosterone acetate have been reported to activate NADPH, trigger oxidative stress and release superoxides [157, 158].

XO is a hypoxia-inducible enzyme, which is found in vascular smooth muscle cells (VSMCs) and vascular endothelial cells. XO catalyzes superoxide production [149]. Mervaala and colleagues [159] documented that rodents with over-expressed human renin and angiotensinogen genes have raised XO activity with endothelial dysfunction and hypertension. Experimental studies have also demonstrated increased renal XO activity in salt-fed spontaneously hypertensive rats [160].

Membrane-bound vascular-derived NOX enzymatic complex can be activated by angiotensin II and aldosterone even at low concentrations [161]. Activation of this system is a major source of ROS, which leads to NO depletion and endothelial dysfunction. Mounting number of studies has implicated the crosstalk between NOX and mitochondria with incident eNOS dysregulation/uncoupling and endothelial dysfunction. It has been established that Ang II stimulates mitochondrial ROS (mtROS) formation and opening of the mitochondrial permeability transition pore (mPTP) with resultant leakage of the generated mtROS to the cytosol [162, 163]. This activates the p38 MAPK and JNK signaling with subsequent activation of NOX [162, 163]. Also, NOX could be activated through cSrc-dependent phosphorylation of p47phox, which is triggered by Ang II. Ang II-dependent NOX causes mitochondrial dysfunction with consequent mtROS formation [164, 165]. This cascade of events leads to a robust accumulation of mtROS formation with cardiovascular implications. mtROS-driven phagocytic NOX activation triggers immune cell infiltration and aggravates Ang II-mediated eNOS uncoupling [166], reduced circulatory NO and endothelial dysfunction.

Although enzymatic superoxide dismutase scavenges the superoxides that are produced in the mitochondria during oxidative phosphorylation, this mechanism may be overwhelmed when ROS generation is exaggerated. This leads to mitochondrial DNA damage and endothelial injury [167].

Oxidative stress and atherosclerosis

Atherosclerosis is the common clinical disorder that results from obesity, IR and DM, and hypertension (Fig. 6). It begins with the formation of atheromatous plaque, which is triggered by endothelial ROS generation and accumulation of LDL in the intima. LDL consists of an ApoB protein molecule, triglycerol, cholesterol and its esters, phospholipids, and vitamin E [168]. The presence of TG enhances the influx and accumulation of LDL in the tunica intima, where it is oxidized by ROS, and picked up by macrophages via scavenger receptor (SR) CD36 to produce foam cells [169,170,171]. Other SRs such as SR-AI, SR-AII, MARCO, and SRCL (class A), SR-BI (class B), CD68 (Class D), LOX-1 (class E), SREC-1 (Class F), and SR-PSOX/CXCL16 (Class G) have been reported [172].

Fig. 6
figure 6

The role of oxidative stress in the pathogenesis of atherosclerosis – early phase (a) and late phase (b). This illustration is culled from the illustration of the role of oxidative stress to adipocytes in atherosclerosis by Otani H [137].

Oxidized LDL (OxLDL) is cytotoxic to atherosclerosis-related cells such as T-cells, macrophages, ECs, and SMCs [173] via OxLDL-derived lipid peroxides and hydroperoxides [174]. High concentrations of OxLDL activates caspase 3 in a Fas-independent manner, thus causing apoptosis with characteristic DNA fragmentation [168]; although, caspases 6, 8 and 9 may also be involved. Besides, OxLDL simultaneously triggers necrosis via ROS [168]. Although OxLDL suppresses nuclear factor-kappa B (NF-kB) in long-term, it activates it in short-term in ECs, SMCs, and macrophages [175,176,177]. OxLDL-induced NF-kB activation is via LOX-1. Binding of OxLDL to LOX-1 promotes the production of O- and H2O2, as well as activation of NF-kB via p38 MAP kinase//P13K/ERK1/2 signaling pathway [178, 179], thus eliciting an inflammation in the endothelial cells. NF-kB regulates the expression of MCP-1, P-selectin, E-selectin, ICAM-1, and VCAM-1 [180].

Recruitment of monocyte-macrophage into the intima is regulated by adhesion molecules, integrins, selectins, and chemokines such as monocyte chemo-attractant protein-1 (MCP-1) [181, 182]. ROS does not just oxidize OxLDL; they also up-regulate MCP-1 and other molecules that are responsible for monocyte-macrophage recruitment. Although, SMCs and ECs synthesize MCP-1, SMCs also moves to the tunica intima from the tunica media, where they differentiate into myofibroblasts and increase in the presence of PDGF [183, 184] and insulin via ROS-dependent phosphorylation of serine residues of IRS-1 [185, 186]. These myofibroblasts are responsible for collagen synthesis, which causes intima thickening.

Enhanced inflammatory response and OS promote apoptosis of the foam cells and necrotic lipid core formation [187, 188]. The necrotic lipid core is covered with a collagen fibre-enriched fibrous cup which is lyzed by activation of matrix metalloproteinases (MMP) in the presence of ROS [189, 190] generating an advanced atheromatous plaque called unstable plaque. This narrows the lumen of the artery and raises the intraluminal pressure.

It has been established that coronary artery disease progresses faster in diabetic patients. There are mounting evidence that suggestive that adiposity and impaired glycaemia promote the development of ischaemic heart disease (IHD) even in non-diabetic individual [191, 192]. Although insulin-treated diabetes has been shown to be an independent predictor of late and repeat coronary revascularization [193], study of Komatsu et al. [194] revealed that IR in the general population treated with percutaneous coronary intervention (PCI) enhanced restenosis due to continuous neointimal growth after the first generation drug eluting stent (DES) implantation. Sasso and his colleagues [192] in a prospective longitudinal observational study demonstrated the role of IR and cytokines in the incident of IHD in normoglycaemic subjects. Findings of their study showed that adiponectin levels were independently associated with restenosis; and HOMA-IR and adiponectin were independently associated with de novo IHD and overall new PCI.

Cardiac metabolic memory

Even when glucose level has been restored, chronic rise in glucose concentration as seen in DM stimulates metabolic alterations that modify tissue homeostasis. This is called metabolic memory [195]. Epigenetics is essential in establishing cardiac metabolic memory. Chronic epigenetic effects like histone and DNA methylations are quite stable and may be inherited as a memory by offspring cells [196]. In addition, maternal nutrition as well as in utero exposure may trigger developmental programming which may also be transferred to progenies, thus triggering disorders like CMD [197]. Hyperglycemia may trigger long-term inflammatory and oxidative stress pathways. This is accompanied by resultant persistent or possibly permanent modifications [198]. Experimental study demonstrates hyperglycemia-dependent ROS as a primary trigger of endothelial glycemic memory [199]. Studies have revealed that hyperglycemia stimulates a reversible increase in the concentrations of IL-6 and decline in histone-3 methylation at the IL-6 promoter in cardiomyocytes [200]. This infers that the raised inflammatory gene expression in cardiomyocytes observed in hyperglycemia is secondary to impaired repressive epigenetic histone modifications [200]. It is not unlikely that hyperglycemia-stimulated mitochondrial dysfunction and apoptosis account for cardiac metabolic memory [200]. Also, hyperglycemia triggers developmental control of insulin-like growth factor-1 (IGF-1) receptor in cardiac muscle cells [201]. Human studies have revealed that a rise in adipose tissue accumulation in obesity is linked to enhanced methylation at the hypoxia-inducible factor 3A (HIF3A) locus in blood cells and adipose tissue, but not in the skin [202]. Metabolic intermediates from catabolism of macromolecules serve as co-factors for chromatin-modifying enzymes [203,204,205,206,207,208,209,210,211,212,213,214].

Conclusion and future perspectives

Oxidative stress, through a complex cascade, is essential in the development of CMD. Increasing evidence has demonstrated that ROS via various pathways triggers systemic inflammation and endothelial cell dysfunction through several mechanisms, such as mitochondrial dysfunction and uncoupling, raised FAO, up-regulation of NOX activity, impaired antioxidant capacity, and cardiac metabolic memory. Although more studies aimed at demonstrating other associated pathogenesis of CMD is important, a good understanding of the link between oxidative stress and CMD opens new therapeutic horizons in the management of CMD by targeting one or more specific pathways in the pathophysiology.

Availability of data and materials

Not applicable

Abbreviations

CMD:

CMD

HDL:

high-density lipoprotein

LDL:

low-density lipoprotein

OxLDL:

Oxidized LDL

TG:

triglycerides

VLDL:

very-low-density lipoprotein

ATP III:

Adult Treatment Panel III

IDF:

International Diabetes Foundation

(NCEP):

National Cholesterol Education Program

TNF-α:

tumour necrosis factor-alpha

IL:

interleukin

MCP-1:

macrophage chemo-attractant protein 1

ROS:

reactive oxygen species

RNS:

reactive nitrogen species

XO:

xanthine oxidase

eNOS:

uncoupled endothelial nitric oxide synthase

HO:

heme oxygenase

XDH:

xanthine dehydrogenase

Ang II:

angiotensin II

GDP:

guanosine diphosphate

PI (3) P:

phosphatidylinositol 3-phosphate

BH4 :

tetrahydrobiopterin

ECs:

endothelial cells

SMCs:

smooth muscle cells

GTP:

guanosine-5’-triphosphate

MAPKs:

mitogen-activated protein kinases

ERK:

extracellular signal-regulated kinases

NO:

nitric oxide

NOX:

nicotinamide adenine dinucleotide phosphate oxidase

O2 - :

superoxide anion radicals

ONOO- :

peroxynitrite

ECM:

extracellular matrix

HIF:

hypoxia-inducible factor

SOD:

superoxide dismutase

H2O2 :

hydrogen peroxide

GPx:

glutathione peroxidase

PON:

paraoxonases

CO:

carbon monoxide

Trx:

thioredoxin

SR:

sarcoplasmic reticulum

SERCA:

sarcoplasmic reticulum ATPase

HNE:

4-hydroxyl 2-nonenal

COX:

cyclooxygenase

RAAS:

renin-angiotensin-aldosterone system

IRS:

insulin receptor substrate

PI3K:

phosphatidylinositol 3-kinase

RBP:

retinol-binding protein

GLUT:

glucose uptake transporter

AT1R:

angiotensin II type 1 receptor

VSMCs:

vascular smooth muscle cells

NF-kB:

nuclear factor-kappa B

LOX:

lipooxygenase

VCAM:

ICAM

PDGF:

platelet-derived growth factor

SMC:

smooth muscle cells

MMP:

matrix metalloproteinases

IGF-1:

insulin-like growth factor-1

FAO:

Fatty acid oxidation

PCI:

percutaneous coronary intervention

DES:

drug eluting stent

References

  1. Grundy SM, Cleeman JI, Daniels SR, Donato KA, Eckel RH, et al. Diagnosis and management of the metabolic syndrome: an American Heart Association/National Heart, Lung, and Blood Institute Scientific Statement. Circulation. 2005;112:2735–52.

    Article  PubMed  Google Scholar 

  2. Han JC, Lawlor DA, Kimm SY. Childhood obesity. Lancet. 2010;375(9727):1737–48.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Malik VS, Willett WC, Hu FB. Global obesity: trends, risk factors and policy implications. Nature Reviews Endocrinol. 2013;9:13–27.

    Article  Google Scholar 

  4. Kuklina EV, Tong X, George MG, Bansil P. Epidemiology and prevention of stroke: a worldwide perspective. Expert Review Neurotherapeutics. 2012;12(2):199–208.

    Article  Google Scholar 

  5. Springer SC, Silverstein J, Copeland K. Management of type 2 diabetes mellitus in children and adolescents. Pediatrics. 2013;131:648–64.

    Article  Google Scholar 

  6. Abete P, Napoli C, Santoro G. Age-related decrease in cardiac tolerance to oxidative stress. J Mol Cell Cardiol. 1999;31(1):227–36.

    Article  CAS  PubMed  Google Scholar 

  7. Brinkley TE, Nicklas BJ, Kanaya AM. Plasma oxidized low-density lipoprotein levels and arterial stiffness in older adults: the health, aging, and body composition study. Hypertension. 2009;53(5):846–52.

    Article  CAS  PubMed  Google Scholar 

  8. Gradinaru D, Borsa C, Ionescu C, Prada GI. Oxidized LDL and NO synthesis – biomarkers of endothelial dysfunction and ageing. Mech Ageing Dev. 2015;151:101–13.

    Article  CAS  PubMed  Google Scholar 

  9. Hossain P, Kawar B, El Nahas M. Obesity and diabetes in the developing world–a growing challenge. N Engl J Med. 2007;356:213–5.

    Article  CAS  PubMed  Google Scholar 

  10. Fezeu L, Balkau B, Kengne AP, Sobngwi E, Mbanya JC. Metabolic syndrome in a sub-Saharan African setting: central obesity may be the key determinant. Atherosclerosis. 2007;193:70–6.

    Article  CAS  PubMed  Google Scholar 

  11. Ulasi II, Ijoma CK, Onodugo OD. A community-based study of hypertension and cardio-metabolic syndrome in semi-urban and rural communities in Nigeria. BMC Health Serv Res. 2010;10:71.

    Article  PubMed  PubMed Central  Google Scholar 

  12. Hosseinpanah F, Barzin M, Sheikholeslami F, Azizi F. Effect of different obesity phenotypes on cardiovascular events in Tehran Lipid and Glucose Study (TLGS). Am J Cardiol. 2011;107:412–6.

    Article  PubMed  Google Scholar 

  13. Harzallah F, Alberti H, Ben KF. The metabolic syndrome in an Arab population: a first look at the new International Diabetes Federation criteria. Diabet Med. 2006;23:441–4.

    Article  CAS  PubMed  Google Scholar 

  14. Abdul-Rahim HF, Husseini A, Bjertness E, Giacaman R, Gordon NH. The metabolic syndrome in the West Bank population: an urban-rural comparison. Diabetes Care. 2001;24:275–9.

    Article  CAS  PubMed  Google Scholar 

  15. Bener A, Zirie M, Musallam M, Khader YS, Al-Hamaq AO. Prevalence of metabolic syndrome according to Adult Treatment Panel III and International Diabetes Federation criteria: a population-based study. Metab Syndr Relat Disord. 2009;7:221–9.

    Article  PubMed  Google Scholar 

  16. Sibai A. Prevalence and correlates of metabolic syndrome in an adult Lebanese population. CVD Prevention and Control. 2008;3:83–90.

    Article  Google Scholar 

  17. Ervin B. Prevalence of Metabolic Syndrome Among Adults 20 Years of Age and Over, by Sex, Age, Race and Ethnicity, and Body Mass Index: United States, 2003–2006. Natl Health Stat Report. 2009;5:1–7.

    Google Scholar 

  18. Grundy SM. Metabolic syndrome pandemic. Arterioscler Thromb Vasc Biol. 2008;28:629–36.

    Article  CAS  PubMed  Google Scholar 

  19. Esposito K, Giugliano F, Martedì E, Feola G, Marfella R, et al. High proportions of erectile dysfunction in men with the metabolic syndrome. Diabetes Care. 2005;28:1201–3.

    Article  PubMed  Google Scholar 

  20. Méndez-Sánchez N, Chavez-Tapia NC, Motola-Kuba D, Sanchez-Lara K, Ponciano-Rodríguez G, et al. Metabolic syndrome as a risk factor for gallstone disease. World J Gastroenterol. 2005;11:1653–7.

    Article  PubMed  PubMed Central  Google Scholar 

  21. Ajayi AF, Akhigbe RE, Ajayi LO. Activation of Cardiac TNF-α In Altered Thyroid State-Induced Cardiometabolic Disorder. J Cardiovasc Disease Res. 2017;8(4):151–6.

    Article  CAS  Google Scholar 

  22. Kelli HM, Kassas I, Lattouf OM. Cardio Metabolic Syndrome: A Global Epidemic. J Diabetes Metab. 2015;6:3.

    Google Scholar 

  23. Castro JP, El-Atat FA, McFarlane SI, Aneja A, Sowers JR. Cardiometabolic syndrome: pathophysiology and treatment. Curr Hypertens Rep. 2003;5:393–401.

    Article  PubMed  Google Scholar 

  24. Petersen KF, Dufour S, Savage DB, Bilz S, Solomon G, et al. The role of skeletal muscle IR in the pathogenesis of the metabolic syndrome. Proc Natl Acad Sci USA. 2007;104:12587–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Abate N, Garg A, Peshock RM, Stray-Gundersen J, Adams-Huet B, Grundy SM. Relationship of generalized and regional adiposity to insulin sensitivity in men with NIDDM. Diabetes. 1996;45:1684–93.

    Article  CAS  PubMed  Google Scholar 

  26. Pouliot MC, Despres JP, Nadeau A, et al. Visceral obesity in men. Associations with glucose tolerance, plasma insulin, and lipoprotein levels. Diabetes. 1992;41:826–34.

    Article  CAS  PubMed  Google Scholar 

  27. Lewis GF, Uffelman KD, Szeto LW, Weller B, Steiner G. Interaction between FFAs and insulin in the acute control of very low density lipoprotein production in humans. J Clin Invest. 1995;95:158–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Zhang YL, Hernandez-Ono A, Ko C, Yasunaga K, Huang LS, Ginsberg HN. Regulation of hepatic apolipoprotein B-lipoprotein assembly and secretion by the availability of fatty acids. I. Differential response to the delivery of fatty acids via albumin or remnant-like emulsion particles. J Biol Chem. 2004;279:19362–74.

    Article  CAS  PubMed  Google Scholar 

  29. Mittendorfer B, Liem O, Patterson BW, Miles JM, Klein S. What does the measurement of wholebody fatty acid rate of appearance in plasma by using a fatty acid tracer really mean? Diabetes. 2003;52:1641–8.

    Article  CAS  PubMed  Google Scholar 

  30. Mittendorfer B, Patterson BW, Klein S. Effect of sex and obesity on basal VLDL-triacylglycerol kinetics. Am J Clin Nutr. 2003;77:573–9.

    Article  CAS  PubMed  Google Scholar 

  31. Steinberg HO, Tarshoby M, Monestel R, et al. Elevated circulating FFA levels impair endothelium-dependent vasodilation. J Clin Invest. 1997;100:1230–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Perseghin G, Petersen K, Shulman GI. Cellular mechanism of IR: potential links with inflammation. Int J Obes Relat Metab Disord. 2003;27(Suppl 3):S6–11.

    Article  CAS  PubMed  Google Scholar 

  33. Crespo J, Cayon A, Fernandez-Gil P, et al. Gene expression of tumor necrosis factor alpha and TNF-receptors, p55 and p75, in nonalcoholic steatohepatitis patients. Hepatology. 2001;34:1158–63.

    Article  CAS  PubMed  Google Scholar 

  34. Park SH, Kim BI, Yun JW, et al. IR and C-reactive protein as independent risk factors for non-alcoholic fatty liver disease in non-obese Asian men. J Gastroenterol Hepatol. 2004;19:694–8.

    Article  CAS  PubMed  Google Scholar 

  35. Baggiolini M, Loetscher P, Moser B. Interleukin-8 and the chemokine family. Int J Immunopharmacol. 1995;17:103–8.

    Article  CAS  PubMed  Google Scholar 

  36. Christiansen T, Richelsen B, Bruun JM. Monocyte chemoattractant protein-1 is produced in isolated adipocytes, associated with adiposity and reduced after weight loss in morbid obese subjects. Int J Obes (Lond). 2005;29:146–50.

    Article  CAS  Google Scholar 

  37. Berg AH, Combs TP, Du X, Brownlee M, Scherer PE. The adipocyte-secreted protein Acrp30 enhances hepatic insulin action. Nat Med. 2001;7:947–53.

    Article  CAS  PubMed  Google Scholar 

  38. Yamauchi T, Kamon J, Minokoshi Y, et al. Adiponectin stimulates glucose utilization and fatty acid oxidation by activating AMP-activated protein kinase. Nat Med. 2002;8:1288–95.

    Article  CAS  PubMed  Google Scholar 

  39. Abate N, Garg A, Peshock RM, Stray-Gundersen J, Grundy SM. Relationships of generalized and regional adiposity to insulin sensitivity in men. J Clin Invest. 1995;96:88–98.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Ross R, Aru J, Freeman J, Hudson R, Janssen I. Abdominal adiposity and IR in obese men. Am J Physiol Endocrinol Metab. 2002;282:E657–63.

    Article  CAS  PubMed  Google Scholar 

  41. Nielsen S, Guo Z, Johnson CM, Hensrud DD, Jensen MD. Splanchnic lipolysis in human obesity. J Clin Invest. 2004;113:1582–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Havel RJ, Kane JP, Balasse EO, Segel N, Basso LV. Splanchnic metabolism of FFAs and production of triglycerides of very low density lipoproteins in normotriglyceridemic and hypertriglyceridemic humans. J Clin Invest. 1970;49:2017–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Davis FM, Gallagher K. Epigenetic Mechanisms in Monocytes/Macrophages Regulate Inflammation in Cardiometabolic and Vascular Disease. Arterioscler Thromb Vasc Biol. 2019;39:00-00. DOI: https://doi.org/10.1161/ATVBAHA.118.312135.

  44. Kirk EP, Klein S. Pathogenesis and Pathophysiology of the Cardiometabolic Syndrome. J Clin Hypertens (Greenwich). 2009;11(12):761–5.

    Article  CAS  Google Scholar 

  45. Apel K, Hirt H. Reactive oxygen species: Metabolism, Oxidative Stress, and Signal Transduction. Annu. Rev. Plant Biol. 2004;55:373–99.

    Article  CAS  PubMed  Google Scholar 

  46. Keshari AK, Farooqi H. Evaluation of the effect of hydrogen peroxide(H2O2) on haemoglobin and the protective effect of glycine” International J Sci Tecnhnoledge. 2014;2(2):36-41.

  47. Gueteens G, De Boeck G, Highley M, Osterom AT, De Bruijn EA. Oxidative DNA damage: Biological significance and methods of analysis. Crit Rev Clin Lab Sci. 2002;39:331–457.

    Article  Google Scholar 

  48. Martin TA, Harrison G, Mansel RE, Jiang WG. The role of the CD44/ezrin complex in cancer metastasis. Crit Rev Oncol Hematol. 2003;46:165–86.

    Article  PubMed  Google Scholar 

  49. Espinosa-Diez C, Miguel V, Mennerich D, Kietzmann T, Sánchez-Pérez P, Cadenas S, Lamas S. Antioxidant responses and cellular adjustments to oxidative stress. Redox Biology. 2015;6:183–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Sack MN, Fyhrquist FY, Saijonmaa OJ, Fuster V, Kovacic JC. Basic Biology of Oxidative Stress and the Cardiovascular System. J Am Coll Cardiol. 2017;70:196–211.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Lee MY, Griendling KK. Redox signaling, vascular function, and hypertension. Antioxidants Redox Signaling. 2008;10(6):1045–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Harris CM, Sanders SA, Massey V. Role of the flavin midpoint potential and NAD binding in determining NAD versus oxygen reactivity of xanthine oxidoreductase. J Biol Chem. 1999;274(8):4561–9.

    Article  CAS  PubMed  Google Scholar 

  53. Landmesser U, Spiekermann S, Preuss C, et al. Angiotensin II induces endothelial xanthine oxidase activation: role for endothelial dysfunction in patients with coronary disease. Arterioscler Thromb Vasc Biol. 2007;27:943–8.

    Article  CAS  PubMed  Google Scholar 

  54. Nauseef WM. Assembly of the phagocyte NADPH oxidase. Histochemistry and Cell Biology. 2004;122(4):277–91.

    Article  CAS  PubMed  Google Scholar 

  55. Ando S, Kaibuchi K, Sasaki T, Hiraoka K, Nishiyama T, Mizuno T, Asada M, Nunoi H, Matsuda I, Matsuura Y. Post-translational processing of rac p21s is important both for their interaction with the GDP/GTP exchange proteins and for their activation of NADPH oxidase. Journal of Biological Chemistry. 1999;267(36):25709–13.

    Article  Google Scholar 

  56. Diebold BA, Bokoch GM. Molecular basis for Rac2 regulation of phagocyte NADPH oxidase. Nature Immunology. 2001;2(3):211–5.

    Article  CAS  PubMed  Google Scholar 

  57. Cave AC, Brewer AC, Narayanapanicker A, Ray R, Grieve DJ, Walker S. NADPH oxidases in cardiovascular health and disease. Antioxidan Redox Signaling. 2006;8(5-6):691–728.

    Article  CAS  Google Scholar 

  58. Ide T, Tsutsui H, Hayashidani S, Kang D, Suematsu N, Nakamura K, Utsumi H, Hamasaki N, Takeshita A. Mitochondrial DNA damage and dysfunction associated with oxidative stress in failing hearts after myocardial infarction. Circulation Research. 2001;88(5):529–35.

    Article  CAS  PubMed  Google Scholar 

  59. Valenti VE, de Abreu LC, Ferreira C and Saldiva PHN. Reactive Oxygen Species and Cardiovascular Diseases, Oxidative Stress and Diseases, Dr. Volodymyr Lushchak (Ed.), InTech, 2012. ISBN: 978-953-51-0552-7, Available from: http://www.intechopen.com/books/oxidative-stress-anddiseases/reactive-oxygen-species-and-cardiovascular-diseases.

  60. Verhaar MC, Westerweel PE, van Zonneveld AJ, Rabelink TJ. Free radical production by dysfunctional eNOS. Heart. 2004;90(5):494–5 ISSN 1468-201X.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Li J. M, Shah A.M. Endothelial cell superoxide generation: regulation and relevance for cardiovascular pathophysiology. Am J Physiol 2004;287(5):R1014–R1030.R1030.

  62. Matsuzawa A, Ichijo H. Stress-responsive protein kinases in redox-regulated apoptosis signaling. Antioxidants and Redox Signaling. 2005;7(3-4):472–81.

    Article  CAS  PubMed  Google Scholar 

  63. Valenti VE, de Abreu LC, Ferreira C and Saldiva PHN. Reactive Oxygen Species and Cardiovascular Diseases, Oxidative Stress and Diseases, Dr. Volodymyr Lushchak (Ed.), 2012, ISBN: 978-953-51-0552-7, InTech, Available from: http://www.intechopen.com/books/oxidative-stress-anddiseases/reactive-oxygen-species-and-cardiovascular-diseases.

  64. Münzel T, Camici GG, Maack C, Bonetti NR, Fuster V, Kovacic JC. Impact of Oxidative Stress on the Heart and Vasculature. J Am College Cardiol. 2017;70(2):212–29.

    Article  CAS  Google Scholar 

  65. Freed JK, Gutterman DD. Mitochondrial reactive oxygen species and vascular function: less is more. Arterioscler Thromb Vasc Biol. 2013;33:673–5.

    Article  CAS  PubMed  Google Scholar 

  66. Stocker R, Keaney JF Jr. Role of oxidative modifications in atherosclerosis. Physiol Rev. 2004;84:1381–478.

    Article  CAS  PubMed  Google Scholar 

  67. Griendling KK, Sorescu D, Lassègue B, Ushio-Fukai M. Modulation of protein kinase activity and gene expression by reactive oxygen species and their role in vascular physiology and pathophysiology. Arterioscler Thromb Vasc Biol. 2000;20:2175–83.

    Article  CAS  PubMed  Google Scholar 

  68. Förstermann U. Nitric oxide and oxidative stress in vascular disease. Pflugers Archiv. 2010;459:923–39.

    Article  PubMed  CAS  Google Scholar 

  69. Förstermann U, Münzel T. Endothelial nitric oxide synthase in vascular disease: from marvel to menace. Circulation. 2006;113:1708–14.

    Article  PubMed  CAS  Google Scholar 

  70. Gryglewski RJ, Palmer RM, Moncada S. Superoxide anion is involved in the breakdown of endothelium-derived vascular relaxing factor. Nature. 1986;320:454–6.

    Article  CAS  PubMed  Google Scholar 

  71. Beckman JS. Oxidative damage and tyrosine nitration from peroxynitrite. Chem Res Toxicol. 1996;9:836–44.

    Article  CAS  PubMed  Google Scholar 

  72. Laursen JB, Somers M, Kurz S, et al. Endothelial regulation of vasomotion in apoE-deficient mice: implications for interactions between peroxynitrite and tetrahydrobiopterin. Circulation. 2001;103:1282–8.

    Article  CAS  PubMed  Google Scholar 

  73. Harrison DG, Chen W, Dikalov S, Li L. Regulation of endothelial cell tetrahydrobiopterin pathophysiological and therapeutic implications. Adv Pharmacol. 2010;60:107–32.

    Article  CAS  PubMed  Google Scholar 

  74. Li H, Horke S, Förstermann U. Vascular oxidative stress, nitric oxide and atherosclerosis. Atherosclerosis. 2014;237:208–19.

    Article  CAS  PubMed  Google Scholar 

  75. Schiffrin EL. Remodeling of resistance arteries in essential hypertension and effects of antihypertensive treatment. Am J Hypertension. 2004;17(12):1192–200.

    Article  CAS  Google Scholar 

  76. Korsgaard N, Aalkjaer C, Heagerty AM, Izzard AS, Mulvany MJ. Histology of subcutaneous small arteries from patients with essential hypertension. Hypertension. 1993;22(4):523–6.

    Article  CAS  PubMed  Google Scholar 

  77. Rizzoni D, Porteri E, Guefi D, Piccoli A, Castellano M, Pasini G, Muiesan ML, Mulvany MJ, Rosei EA. Cellular hypertrophy in subcutaneous small arteries of patients with renovascular hypertension. Hypertension. 2000;35(4):931–5.

    Article  CAS  PubMed  Google Scholar 

  78. Tribble DL, Gong EL, Leeuwenburgh C. Fatty streak formation in fat-fed mice expressing human copper-zinc superoxide dismutase. Arterioscler Thromb Vasc Biol. 1997;17:1734–40.

    Article  CAS  PubMed  Google Scholar 

  79. Sentman ML, Brännstrom T, Westerlund S. Extracellular superoxide dismutase deficiency and atherosclerosis in mice. Arterioscler Thromb Vasc Biol. 2001;21:1477–82.

    Article  CAS  PubMed  Google Scholar 

  80. Yang H, Roberts LJ, Shi MJ. Retardation of atherosclerosis by overexpression of catalase or both Cu/Zn-superoxide dismutase and catalase in mice lacking apolipoprotein E. Circ Res. 2004;95:1075–81.

    Article  CAS  PubMed  Google Scholar 

  81. Fukai T, Ushio-Fukai M. Superoxide dismutases: role in redox signaling, vascular function, and diseases. Antioxid Redox Signal. 2011;15:1583–606.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Zhang Y, Griendling KK, Dikalova A, Owens GK, Taylor WR. Vascular hypertrophy in angiotensin II-induced hypertension is mediated by vascular smooth muscle cell-derived H2O2. Hypertension. 2005;46:732–7.

    Article  CAS  PubMed  Google Scholar 

  83. Lewis P, Stefanovic N, Pete J. Lack of the antioxidant enzyme glutathione peroxidase-1 accelerates atherosclerosis in diabetic apolipoprotein E-deficient mice. Circulation. 2007;115:2178–87.

    Article  CAS  PubMed  Google Scholar 

  84. Torzewski M, Ochsenhirt V, Kleschyov AL. Deficiency of glutathione peroxidase-1 accelerates the progression of atherosclerosis in apolipoprotein E-deficient mice. Arterioscler Thromb Vasc Biol. 2007;27:850–7.

    Article  CAS  PubMed  Google Scholar 

  85. Yoshida T, Maulik N, Engelman RM. Glutathione peroxidase knockout mice are susceptible to myocardial ischemia reperfusion injury. Circulation. 1997;96:II216–20.

    Google Scholar 

  86. Blankenberg S, Rupprecht HJ, Bickel C. AtheroGene Investigators. Glutathione peroxidase 1 activity and cardiovascular events in patients with coronary artery disease. N Engl J Med. 2003;349:1605–13.

    Article  CAS  PubMed  Google Scholar 

  87. Guo Z, Ran Q, Roberts LJ. Suppression of atherogenesis by overexpression of glutathione peroxidase-4 in apolipoprotein E-deficient mice. Free Radic Biol Med. 2008;44:343–52.

    Article  CAS  PubMed  Google Scholar 

  88. Witte I, Foerstermann U, Devarajan A, Reddy ST, Horke S. Protectors or traitors: the roles of PON2 and PON3 in atherosclerosis and cancer. J Lipids. 2012;2012:342806.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  89. Huang Y, Wu Z, Riwanto M. Myeloperoxidase, paraoxonase-1, and HDL form a functional ternary complex. J Clin Invest. 2013;123:3815–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Horke S, Witte I, Wilgenbus P, Krüger M, Strand D, Förstermann U. Paraoxonase-2 reduces oxidative stress in vascular cells and decreases endoplasmic reticulum stress-induced caspase activation. Circulation. 2007;115:2055–64.

    Article  CAS  PubMed  Google Scholar 

  91. Ng CJ, Bourquard N, Grijalva V. Paraoxonase-2 deficiency aggravates atherosclerosis in mice despite lower apolipoprotein-B containing lipoproteins: anti-atherogenic role for paraoxonase-2. J Biol Chem. 2006;281:29491–500.

    Article  CAS  PubMed  Google Scholar 

  92. Schweikert EM, Devarajan A, Witte I. PON3 is upregulated in cancer tissues and protects against mitochondrial superoxide-mediated cell death. Cell Death Differ. 2012;19:1549–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Shih DM, Xia YR, Wang XP. Decreased obesity and atherosclerosis in human paraoxonase 3 transgenic mice. Circ Res. 2007;100:1200–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Marsillach J, Camps J, Beltran-Debón R. Immunohistochemical analysis of paraoxonases-1 and 3 in human atheromatous plaques. Eur J Clin Invest. 2011;41:308–14.

    Article  CAS  PubMed  Google Scholar 

  95. Stocker R, Perrella MA. Heme oxygenase-1: a novel drug target for atherosclerotic diseases? Circulation. 2006;114:2178–89.

    Article  CAS  PubMed  Google Scholar 

  96. Taille C, El-Benna J, Lanone S. Induction of heme oxygenase-1 inhibits NAD(P) H oxidase activity by down-regulating cytochrome b558 expression via the reduction of heme availability. J Biol Chem. 2004;279:28681–8.

    Article  CAS  PubMed  Google Scholar 

  97. Jiang F, Roberts SJ. Datla Sr., Dusting GJ. NO modulates NADPH oxidase function via heme oxygenase-1 in human endothelial cells. Hypertension. 2006;48:950–7.

    Article  CAS  PubMed  Google Scholar 

  98. Hilgers RH, Kundumani-Sridharan V, Subramani J. Thioredoxin reverses age related hypertension by chronically improving vascular redox and restoring eNOS function. Sci Transl Med. 2017;9:eaaf6094.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  99. May JM. How does ascorbic acid prevent endothelial dysfunction? Free Radic Biol Med. 2000;28:1421–9.

    Article  CAS  PubMed  Google Scholar 

  100. Heller R, Unbehaun A, Schellenberg B, Mayer B, Werner-Felmayer G, Werner ER. L-Ascorbic acid potentiates endothelial nitric oxide synthesis via a chemical stabilization of tetrahydrobiopterin. J Biol Chem. 2001;276:40–7.

    Article  CAS  PubMed  Google Scholar 

  101. Wallerath T, Deckert G, Ternes T. Resveratrol, a polyphenolic phytoalexin present in red wine, enhances expression and activity of endothelial nitric oxide synthase. Circulation. 2002;106:1652–8.

    Article  CAS  PubMed  Google Scholar 

  102. Wallerath T, Li H, Gödtel-Ambrust U, Schwarz PM, Förstermann U. A blend of polyphenolic compounds explains the stimulatory effect of red wine on human endothelial NO synthase. Nitric Oxide. 2005;12:97–104.

    Article  CAS  PubMed  Google Scholar 

  103. Agarwal A, Sekhon LH. Oxidative stress and antioxidants for idiopathic oligoasthenoteratospermia: Is it justified? Indian J Urol. 2011;27:74–85.

    Article  PubMed  PubMed Central  Google Scholar 

  104. Agarwal A, Saleh RA. Role of oxidants in male infertility: rationale, significance, and treatment. Urol Clin North Am. 2002;29:817–27.

    Article  PubMed  Google Scholar 

  105. Griveau JF, Dumont E, Renard B, Callegari JP, Lannou D. Reactive oxygen species, lipid peroxidation and enzymatic defense systems in human spermatozoa. J Reprod Fertil. 1995;103:17–26.

    Article  CAS  PubMed  Google Scholar 

  106. Hool L. C, Corry B. Redox control of calcium channels: from mechanisms to therapeutic opportunities. Antioxidant and Redox Signal 2007; 4:409–435.

  107. Cherednichenko G, Zima AV, Feng W, Schaefer S, Blatter LA, Pessah IN. NADH oxidase activity of rat cardiac sarcoplasmic reticulum regulates calcium induced calcium release. Circulation Research. 2004;94(4):478–86.

    Article  CAS  PubMed  Google Scholar 

  108. Sanchez G, Escobar M, Pedrozo Z, Macho P, Domenech R, Hartel S. Exercise and tachycardia increase NADPH oxidase and ryanodine receptor-2 activity: possible role in cardioprotection. Cardiovascular Research. 2008;77(2):380–6.

    Article  CAS  PubMed  Google Scholar 

  109. Yi XY, Li VX, Zhang F, Yi F, Matson DR, Jiang MT. Characteristics and actions of NAD(P) H oxidase on the sarcoplasmic reticulum of coronary artery smooth muscle. Am J Physiol Heart Circulatory Physiol. 2006;290(3):H1136–44.

    Article  CAS  Google Scholar 

  110. Zeng Q, Zhou Q, Yao F, O’Rourke S. T, Sun C. Endothelin-1 regulates cardiac Ltype calcium channels via NAD(P) H oxidase-derived superoxide. J Pharmacol Experimental Therapeutics 2006; 326(3): 732–738.

  111. Montague CT, O’Rahilly S. The perils of portliness: causes and consequences of visceral adiposity. Diabetes. 2000;49:883–8.

    Article  CAS  PubMed  Google Scholar 

  112. Matsuzawa Y, Funahashi T, Nakamura T. Molecular mechanism of metabolic syndrome X: Contribution of adipocytokines adipocyte-derived bioactive substances. Ann. N. Y.Acad Sci. 1999;892:146–54.

    Article  CAS  PubMed  Google Scholar 

  113. Spiegelman BM, Flier JS. Obesity and the regulation of energy balance. Cell. 2001;104:531–43.

    Article  CAS  PubMed  Google Scholar 

  114. Kahn BB, Flier JS. Obesity and IR. J. Clin. Invest. 2000;106:473–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Furukawa S, Fujita T, Shimabukuro M, Iwaki M, Yamada Y, Nakajima Y, Nakayama O, Makishima M, Matsuda M, Shimomura I. Increased oxidative stress in obesity and its impact on metabolic syndrome. J Clin Invest. 2004;114:1752–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Devaraj S, Wang-Polagruto J, Polagruto J, Keen CL, Jialal I. High-fat, energy-dense, fast-food-style breakfast results in an increase in oxidative stress in metabolic syndrome. Metab Clin Exp. 2008;57:867–70.

    Article  CAS  PubMed  Google Scholar 

  117. Johnson JB, Summer W, Cutler RG, Martin B, Hyun D-H, Dixit VD, Pearson M, Nassar M, Telljohann R, Tellejohan R, Maudsley S, Carlson O, John S, Laub DR, Mattson MP. Alternate day calorie restriction improves clinical findings and reduces markers of oxidative stress and inflammation in overweight adults with moderate asthma. Free Radic Biol Med. 2007;42:665–74.

    Article  CAS  PubMed  Google Scholar 

  118. Hattori Y, Akimoto K, Gross SS, Hattori S, Kasai K. Angiotensin-II-induced oxidative stress elicits hypoadiponectinaemia in rats. Diabetologia. 2005;48:1066–74.

    Article  CAS  PubMed  Google Scholar 

  119. Soares AF, Guichardant M, Cozzone D, Bernoud-Hubac N, Bouzaidi-Tiali N, Lagarde M, Geloen A. Effects of oxidative stress on adiponectin secretion and lactate production in 3T3-L1 adipocytes. Free Radic Biol Med. 2005;38:882–9.

    Article  CAS  PubMed  Google Scholar 

  120. Chen B, Wei J, Wang W, Cui G, Zhao Y, Zhu X, Zhu M, Guo W, Yu J. Identification of signaling pathways involved in aberrant production of adipokines in adipocytes undergoing oxidative stress. Arch Med Res. 2009;40:241–8.

    Article  CAS  PubMed  Google Scholar 

  121. Sakurai T, Izawa T, Kizaki T, Ogasawara JE, Shirato K, Imaizumi K, Takahashi K, Ishida H, Ohno H. Exercise training decreases expression of inflammation-related adipokines through reduction of oxidative stress in rat white adipose tissue. Biochem Biophys Res Commun. 2009;379:605–9.

    Article  CAS  PubMed  Google Scholar 

  122. Shimomura I. Enhanced expression of PAI-1 in visceral fat: possible contributor to vascular disease in obesity. Nat. Med. 1996;2:800–3.

    Article  CAS  PubMed  Google Scholar 

  123. Hotamisligil GS, Shargill NS, Spiegelman BM. Adipose expression of tumor necrosis factor-α: direct role in obesity-linked IR. Science. 1993;259:87–91.

    Article  CAS  PubMed  Google Scholar 

  124. Uysal KT, Wiesbrock SM, Marino MW, Hotamisligil GS. Protection from obesity-induced IR in mice lacking TNF-alpha function. Nature. 1997;389:610–4.

    Article  CAS  PubMed  Google Scholar 

  125. Fruebis J, et al. Proteolytic cleavage product of 30-kDa adipocyte complement-related protein increases fatty acid oxidation in muscle and causes weight loss in mice. Proc. Natl. Acad. Sci. U. S. A. 2001;98:2005–10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Yamauchi T, et al. The fat-derived hormone adiponectin reverses IR associated with both lipoatrophy and obesity. Nat. Med. 2001;7:941–6.

    Article  CAS  PubMed  Google Scholar 

  127. Maeda N, et al. Diet-induced IR in mice lacking adiponectin/ACRP30. Nat. Med. 2002;8:731–7.

    Article  CAS  PubMed  Google Scholar 

  128. Okamoto Y, et al. Adiponectin reduces atherosclerosis in apolipoprotein E-deficient mice. Circulation. 2002;106:2767–70.

    Article  CAS  PubMed  Google Scholar 

  129. Matsuda M, et al. Role of adiponectin in preventing vascular stenosis. The missing link of adipovascular axis. J. Biol. Chem. 2002;277:37487–91.

    Article  CAS  PubMed  Google Scholar 

  130. Yamauchi T, et al. Globular adiponectin protected ob/ob mice from diabetes and apoE-deficient mice from atherosclerosis. J. Biol. Chem. 2003;278:2461–8.

    Article  CAS  PubMed  Google Scholar 

  131. Zarrouki B, Soares AF, Guichardant M, Lagarde M, Geloen A. The lipid peroxidation end-product 4-HNE induces COX-2 expression through p38MAPK activation in 3T3-L1 adipose cell. FEBS Lett. 2007;581:2394–400.

    Article  CAS  PubMed  Google Scholar 

  132. Wang Z, Dou X, Gu D, Shen C, Yao T, Nguyen V, Braunschweig C, Song Z. 4- Hydroxynonenal differentially regulates adiponectin gene expression and secretion via activating PPARγ and accelerating ubiquitin-proteasome degradation. Mol Cell Endocrinol. 2012;349:222–31.

    Article  CAS  PubMed  Google Scholar 

  133. Cassis LA, Police SB, Yiannikouris F, Thatcher SE. Local adipose tissue renin-angiotensin system. Curr Hypertens Rep. 2008;10:93–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Dandona P, Dhindsa S, Ghanim H, Chaudhuri A. Angiotensin II and inflammation: the effect of angiotensinconverting enzyme inhibition and angiotensin II receptor blockade. J Hum Hypertens. 2007;21:20–7.

    Article  CAS  PubMed  Google Scholar 

  135. Koh KK, Oh PC, Quon MJ. Does reversal of oxidative stress and inflammation provide vascular protection? Cardiovasc Res. 2009;81:649–59.

    Article  CAS  PubMed  Google Scholar 

  136. Skultetyova D, Filipova S, Riecansky I, Skultety J. The role of angiotensin type 1 receptor in inflammation and endothelial dysfunction. Recent Pat Cardiovasc Drug Discov. 2007;2:23–7.

    Article  CAS  PubMed  Google Scholar 

  137. Otani H. Oxidative Stress as Pathogenesis of Cardiovascular Risk Associated with Metabolic Syndrome. Antioxidants Redox Signaling. 2011;15(7):1911–26.

    Article  CAS  PubMed  Google Scholar 

  138. Bendall JK, Alp NJ, Warrick N, Cai S, Adlam D, Rockett K, Yokoyama M, Kawashima S, Channon KM. Stoichiometric relationships between endothelial tetrahydrobiopterin, endothelial NO synthase (eNOS) activity, and eNOS coupling in vivo: insights from transgenic mice with endothelial-targeted GTP cyclohydrolase 1 and eNOS overexpression. Circ Res. 2005;97:864–71.

    Article  CAS  PubMed  Google Scholar 

  139. Schulz E, Jansen T, Wenzel P, Daiber A, Munzel T. Nitric oxide, tetrahydrobiopterin, oxidative stress, and endothelial dysfunction in hypertension. Antioxid Redox Signal. 2008;10:1115–26.

    Article  CAS  PubMed  Google Scholar 

  140. Akhigbe R, Ajayi A. Testicular toxicity following chronic codeine administration is via oxidative DNA damage and up-regulation of NO/TNF-α and caspase 3 activities. PLoS ONE. 2020;15(3):e0224052.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Dickhout JG, Hossain GS, Pozza LM, Zhou J, Lhotak S, Austin RC. Peroxynitrite causes endoplasmic reticulum stress and apoptosis in human vascular endothelium: implications in atherogenesis. Arterioscler Thromb Vasc Biol. 2005;25:2623–9.

    Article  CAS  PubMed  Google Scholar 

  142. Duda DG, Fukumura D, Jain RK. Role of eNOS in neovascularization: NO for endothelial progenitor cells. Trends Mol Med. 2004;10:143–5.

    Article  CAS  PubMed  Google Scholar 

  143. Luque Contreras D, Vargas Robles H, Romo E, Rios A, Escalante B. The role of nitric oxide in the post-ischemic revascularization process. Pharmacol Ther. 2006;112:553–63.

    Article  CAS  PubMed  Google Scholar 

  144. Archuleta TL, Lemieux AM, Saengsirisuwan V, Teachey MK, Lindborg KA, Kim JS, Henriksen EJ. Oxidant stress-induced loss of IRS-1 and IRS-2 proteins in rat skeletal muscle: role of p38 MAPK. Free Radic Biol Med. 2009;47:1486–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Tamori Y, Sakaue H, Kasuga M. RBP4, an unexpected adipokine. Nat Med. 2006;12:30–1; discussion 31.

    Article  CAS  PubMed  Google Scholar 

  146. Muoio DM, Newgard CB. Metabolism: A is for adipokine. Nature. 2005;436:337–8.

    Article  CAS  PubMed  Google Scholar 

  147. Beckman JS, Crow JP. Pathological implications of nitric oxide, superoxide and peroxynitrite formation. Biochem. Soc. Trans. 1993;21:330–4.

    Article  CAS  PubMed  Google Scholar 

  148. Beckman JS, Beckman TW, Chen J. Apparent hydroxyl radical production of peroxynitrite: implications for endothelial injury from nitric oxide and superoxide. Proc. Natl. Acad. Sci. USA. 1990;87:1620–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Manrique C, Lastra G, Gardner M, Sowers JR. The Renin Angiotensin Aldosterone System in Hypertension: Roles of IR and Oxidative Stress. Med Clin North Am. 2009;93(3):569–82. https://doi.org/10.1016/j.mcna.2009.02.014.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Madamanchi NR, Vendrov A, Runge MS. Oxidative stress and vascular disease. Arterioscler Thromb Vasc Biol. 2005;25:29–38.

    Article  CAS  PubMed  Google Scholar 

  151. Sydow K, Munzel T. ADMA and oxidative stress. Atheroscler Suppl. 2003;4:41–51.

    Article  CAS  PubMed  Google Scholar 

  152. Böger RH, Sydow K, Borlak J, Thum T, Lenzen H, Schubert B, Tsikas D, Bode-Böger SM. LDL cholesterol upregulates synthesis of asymmetrical dimethylarginine in human endothelial cells: involvement of S-adenosylmethionine-dependent methyltransferases. Circ Res. 2000;87:99–105.

    Article  PubMed  Google Scholar 

  153. Zou MH, Shi C, Cohen RA. Oxidation of the zinc-thiolate complex and uncoupling of endothelial nitric oxide synthase by peroxynitrite. J Clin Invest. 2002;109:817–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Massiera F, Bloch-Faure M, Ceiler D, et al. Adipose angiotensinogen is involved in adipose tissue growth and blood pressure regulation. FASEB J. 2001;15:2727–9.

    Article  CAS  PubMed  Google Scholar 

  155. Cooper SA, Whaley-Connell A, Habibi J, et al. Renin-angiotensin-aldosterone system and oxidative stress in cardiovascular IR. Am J Physiol Heart Circ Physiol. 2007;293:H2009–23.

    Article  CAS  PubMed  Google Scholar 

  156. Mehta PK, Griendling KK. Angiotensin II cell signaling: physiological and pathological effects in the cardiovascular system. Am J Physiol, Cell Physiol. 2007;292:C82–97.

    Article  CAS  Google Scholar 

  157. Johar S, Cave AC, Narayanapanicker A, et al. Aldosterone mediates angiotensin II-induced interstitial cardiac fibrosis via a Nox2-containing NADPH oxidase. FASEB J. 2006;20:1546–8.

    Article  CAS  PubMed  Google Scholar 

  158. Beswick RA, Zhang H, Marable D, et al. Long-term antioxidant administration attenuates mineralocorticoid hypertension and renal inflammatory response. Hypertension. 2001;37:781–6.

    Article  CAS  PubMed  Google Scholar 

  159. CS, Cheng ZJ, Tikkanen I, et al. Endothelial dysfunction and xanthine oxidoreductase activity in rats with human renin and angiotensinogen genes. Hypertension. 2001;37:414–8.

    Article  Google Scholar 

  160. Laakso J, Mervaala E, Himberg JJ, et al. Increased kidney xanthine oxidoreductase activity in saltinduced experimental hypertension. Hypertension. 1998;32:902–6.

    Article  CAS  PubMed  Google Scholar 

  161. Griendling KK, Minieri CA, Ollerenshaw JD, et al. Angiotensin II stimulates NADH and NADPH oxidase activity in cultured vascular smooth muscle cells. Circ Res. 1994;74:1141–8.

    Article  CAS  PubMed  Google Scholar 

  162. Brandes RP. Triggering mitochondrial radical release: a new function for NADPH oxidases. Hypertension. 2005;45:847–8.

    Article  CAS  PubMed  Google Scholar 

  163. Kimura S, Zhang GX, Nishiyama A, Shokoji T, Yao L, Fan YY, Rahman M, Abe Y. Mitochondria-derived reactive oxygen species and vascular MAP kinases: comparison of angiotensin II and diazoxide. Hypertension. 2005;45:438–44.

    Article  CAS  PubMed  Google Scholar 

  164. Reinehr R, Becker S, Eberle A, Grether-Beck S, Haussinger D. Involvement of NADPH oxidase isoforms and Src family kinases in CD95-dependent hepatocyte apoptosis. J Biol Chem. 2005;280:27179–94.

    Article  CAS  PubMed  Google Scholar 

  165. Touyz RM, Yao G, Schiffrin EL. c-Src induces phosphorylation and translocation of p47phox: role in superoxide generation by angiotensin II in human vascular smooth muscle cells. Arterioscler Thromb Vasc Biol. 2003;23:981–7.

    Article  CAS  PubMed  Google Scholar 

  166. Kroller-Schon S, Steven S, Kossmann S, Scholz A, Daub S, Oelze M, Xia N, Hausding M, Mikhed Y, Zinßius E, Mader M, Stamm P, Treiber N, Scharffetter-Kochanek K, Li H, Schulz E, Wenzel P, Munzel T, Daiber A. Molecular Mechanisms of the Crosstalk Between Mitochondria and NADPH Oxidase Through Reactive Oxygen Species—Studies in White Blood Cells and in Animal Models. Antioxid. Redox Signal. 2014;20:247–66.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  167. Ballinger SW, Patterson C, Knight-Lozano CA, et al. Mitochondrial integrity and function in atherogenesis. Circulation. 2002;106:544–9.

    Article  CAS  PubMed  Google Scholar 

  168. Garelnabi M, Kakumanu S, Litvinov D. Role of Oxidized Lipids in Atherosclerosis. In: Oxidative Stress and Diseases. Ed: Dr. Volodymyr Lushchak. ISBN: 978-953-51-0552-7. InTech. Available from: http://www.intechopen.com/books/oxidative-stress-and-diseases/role-of-oxidized-lipids-in-atherosclerosis.

  169. Collot-Teixeira S, Martin J, McDermott-Roe C, Poston R, McGregor JL. CD36 and macrophages in atherosclerosis. Cardiovasc Res. 2007;75:468–77.

    Article  CAS  PubMed  Google Scholar 

  170. Rahaman SO, Lennon DJ, Febbraio M, Podrez EA, Hazen SL, Silverstein RL. A CD36-dependent signaling cascade is necessary for macrophage foam cell formation. Cell Metab. 2006;4:211–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Thorne RF, Mhaidat NM, Ralston KJ, Burns GF. CD36 is a receptor for oxidized high density lipoprotein: implications for the development of atherosclerosis. FEBS Lett. 2007;581:1227–32.

    Article  CAS  PubMed  Google Scholar 

  172. Moore KJ, Freeman MW. Scavenger receptors in atherosclerosis: beyond lipid uptake. Arterioscler Thromb Vasc Biol. 2006;26(8):1702–11.

    Article  CAS  PubMed  Google Scholar 

  173. Alcouffe J, Caspar-Bauguil S, Garcia V, Salvayre R, Thomsen M, Benoist H. Oxidized low density lipoproteins induce apoptosis in PHA-activated peripheral blood mononuclear cells and in the Jurkat T-cell line. J Lipid Res. 1999;40(7):1200–10.

    Article  CAS  PubMed  Google Scholar 

  174. Coffey MD, Cole RA, Colles SM, Chisolm GM. In vitro cell injury by oxidized low density lipoprotein involves lipid hydroperoxide-induced formation of alkoxyl, lipid, and peroxyl radicals. J Clin Invest. 1995;96:1866–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Brand K, Eisele T, Kreusel U, Page M, Page S, Haas M, Gerling A, Kaltschmidt C, Neumann FJ, Mackman N, Baeurele PA, Walli AK, Neumeier D. Dysregulation of monocytic nuclear factor-kappa B by oxidized low-density lipoprotein. Arterioscler Thromb Vasc Biol. 1997;17:1901–9.

    Article  CAS  PubMed  Google Scholar 

  176. Eligini S, Brambilla M, Banfi C, Camera M, Sironi L, Barbieri SS, Auwerx J, Tremoli E. Colli S. Oxidized phospholipids inhibit cyclooxygenase-2 in human macrophages via nuclear factor-kappaB/IkappaB-and ERK2-dependent mechanisms. Cardiovasc Res. 2002;55:406–15.

    Article  CAS  PubMed  Google Scholar 

  177. Brand K, Page S, Rogler G, Bartsch A, Brandl R, Knuechel R, Page M, Kaltschmidt C, Baeuerle PA, Neumeier D. Activated transcription factor nuclear factor kappa B is present in the atherosclerotic lesion. J Clin Invest. 1996;97(7):1715–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Cominacini L, Pasini AF, Garbin U, Davoli A, Tosetti ML, Campagnola M, Rigoni A, Pastorino AM, Lo Cascio V, Sawamura T. Oxidized low density lipoprotein (OxLDL) binding to OxLDL receptor-1 in endothelial cells induces the activation of NF-kappaB through an increased production of intracellular reactive oxygen species. J Biol Chem. 2000;275(17):12633–8.

    Article  CAS  PubMed  Google Scholar 

  179. Tanigawa H, Miura S, Matsuo Y, Fujino M, Sawamura T, Saku K. Dominant-negative lox-1 blocks homodimerization of wild-type lox-1-induced cell proliferation through extracellular signal regulated kinase 1/2 activation. Hypertension. 2006;48(2):294–300.

    Article  CAS  PubMed  Google Scholar 

  180. Cominacini L, Garbin U, Fratta Pasini A, Paulon T, Davoli A, Campagnola M, Marchi E, Pastorino AM, Gaviraghi G, Lo CV. Lacidipine inhibits the activation of the transcription factor NF-kappaB and the expression of adhesion molecules induced by pro-oxidant signals on endothelial cells. J Hypertens. 1997;15(12 Pt 2):1633–40.

    Article  CAS  PubMed  Google Scholar 

  181. Wang G, Woo CW, Sung FL, Siow YL, and O K. Increased monocyte adhesion to aortic endothelium in rats with hyperhomocysteinemia: role of chemokine and adhesion molecules. Arterioscler Thromb Vasc Biol 22: 1777–1783, 2002.

  182. Boyle JJ. Macrophage activation in atherosclerosis: pathogenesis and pharmacology of plaque rupture. Curr Vasc Pharmacol. 2005;3:63–8.

    Article  CAS  PubMed  Google Scholar 

  183. DiCorleto PE. Cellular mechanisms of atherogenesis. Am J Hypertens. 1993;6:314S–8S.

    Article  CAS  PubMed  Google Scholar 

  184. Myllarniemi M, Calderon L, Lemstrom K, Buchdunger E, Hayry P. Inhibition of platelet-derived growth factor receptor tyrosine kinase inhibits vascular smooth muscle cell migration and proliferation. FASEB J. 1997;11:1119–26.

    Article  CAS  PubMed  Google Scholar 

  185. Vinayagamoorthi R, Bobby Z, Sridhar MG. Antioxidants preserve redox balance and inhibit c-Jun-Nterminal kinase pathway while improving insulin signaling in fat-fed rats: evidence for the role of oxidative stress on IRS-1 serine phosphorylation and IR. J Endocrinol. 2008;197:287–96.

    Article  CAS  PubMed  Google Scholar 

  186. Igarashi M, Hirata A, Yamaguchi H, Tsuchiya H, Ohnuma H, Tominaga M, Daimon M, Kato T. Candesartan inhibits carotid intimal thickening and ameliorates IR in balloon-injured diabetic rats. Hypertension. 2001;38:1255–9.

    Article  CAS  PubMed  Google Scholar 

  187. Martinet W, Kockx MM. Apoptosis in atherosclerosis:focus on oxidized lipids and inflammation. Curr Opin Lipidol. 2001;12:535–41.

    Article  CAS  PubMed  Google Scholar 

  188. Hung YC, Hong MY, Huang GS. Cholesterol loading augments oxidative stress in macrophages. FEBS Lett. 2006;580:849–61.

    Article  CAS  PubMed  Google Scholar 

  189. Wainwright CL. Matrix metalloproteinases, oxidative stress and the acute response to acute myocardial ischaemia and reperfusion. Curr Opin Pharmacol. 2004;4:132–8.

    Article  CAS  PubMed  Google Scholar 

  190. Libby P. The molecular mechanisms of the thrombotic complications of atherosclerosis. J Intern Med. 2008;263:517–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Levitan EB, Song Y, Ford ES, Liu S. Is nondiabetic hyperglycemia a risk factor for cardiovascular disease? A meta-analysis of prospective studies. Arch Intern Med. 2004;164(19):2147–55 https://doi.org/10.1001/archinte.164.19.2147.

    Article  PubMed  Google Scholar 

  192. Sasso FC, Pafundi PC, Marfella R, Calabrò P, Piscione F, Furbatto F, Esposito G, Galiero R, Gragnano F, Rinald L, Salvatore T, D’Amico M, Adinolfi LE, Sardu S. Adiponectin and insulin resistance are related to restenosis and overall new PCI in subjects with normal glucose tolerance: the prospective AIRE Study. Cardiovasc Diabetol. 2019;18:24 https://doi.org/10.1186/s12933-019-0826-0.

    Article  PubMed  PubMed Central  Google Scholar 

  193. Orbach A, Halon DA, Jaffe R, Rubinshtein R, Karkabi B, Flugelman MY, Zafrir B. Impact of diabetes and early revascularization on the need for late and repeat procedures. Cardiovasc Diabetol. 2018;17(1):25. https://doi.org/10.1186/s12933-018-0669-0.

  194. Komatsu T, Komatsu S, Nakamura H, Kuroyanagi T, Fujikake A, Hisauchi I, Sakuma M, Nakahara S, Sakai Y, Taguchi I. Insulin resistance as a predictor of the late catch-up phenomenon after drug-eluting stent implantation. Circ J. 2016;80(3):657–62. https://doi.org/10.1253/circj.CJ-15-1012

  195. Intine RV, Sarras MP Jr. Metabolic memory and chronic diabetes complications: potential role for epigenetic mechanisms. Curr Diab Rep. 2012;12:551–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  196. Gillette TG, Hill JA. Readers, writers, and erasers: chromatin as the whiteboard of heart disease. Circ Res. 2015;116:1245–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  197. Barnes SK, Ozanne SE. Pathways linking the early environment to long-term health and lifespan. Prog Biophys Mol Biol. 2011;106:323–36.

    Article  CAS  PubMed  Google Scholar 

  198. Wegner M, Neddermann D, Piorunska-Stolzmann M, Jagodzinski PP. Role of epigenetic mechanisms in the development of chronic complications of diabetes. Diabetes Res Clin Pract. 2014;105:164–75.

    Article  CAS  PubMed  Google Scholar 

  199. Keating ST, Plutzky J, El-Osta A. Epigenetic changes in diabetes and cardiovascular risk. Circ Res. 2016;118:1706–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  200. Yu XY, Geng YJ, Liang JL. High levels of glucose induce “metabolic memory” in cardiomyocyte via epigenetic histone H3 lysine 9 methylation. Mol Biol Rep. 2012;39:8891–8.

    Article  CAS  PubMed  Google Scholar 

  201. Yu XY, Geng YJ, Liang JL. High levels of glucose induce apoptosis in cardiomyocyte via epigenetic regulation of the insulin-like growth factor receptor. Exp Cell Res. 2010;316:2903–9.

    Article  CAS  PubMed  Google Scholar 

  202. Dick KJ, Nelson CP, Tsaprouni L. DNA methylation and body-mass index: a genome-wide analysis. Lancet. 2014;383:1990–8.

    Article  CAS  PubMed  Google Scholar 

  203. Keating ST, El-Osta A. Epigenetics and metabolism. Circ Res. 2015;116:715–36.

    Article  CAS  PubMed  Google Scholar 

  204. Scholze J, Alegria E, Ferri C, Langham S, Stevens W, et al. Epidemiological and economic burden of metabolic syndrome and its consequences in patients with hypertension in Germany, Spain and Italy; a prevalence-based model. BMC Public Health. 2010;10:529.

    Article  PubMed  PubMed Central  Google Scholar 

  205. Vancampfort D, Hallgren M, Mugisha J, Hert M, Probst M, Monsieur D, Stubbs B. The Prevalence of Metabolic Syndrome in Alcohol Use Disorders: A Systematic Review and Meta-analysis Alcohol and Alcoholism. 2016; 51(5):515–21.

  206. Goodwin RD, Kim JH, Weinberger AH, et al. Symptoms of alcohol dependence and smoking initiation and persistence: a longitudinal study among US adults. Drug Alcohol Depend. 2013;133:718–23.

    Article  PubMed  PubMed Central  Google Scholar 

  207. Smothers B, Bertolucci D. Alcohol consumption and healthpromoting behavior in a US household sample: leisure-time physical activity. J Stud Alcohol. 2001;62:467–76.

  208. Whang W, Kubzansky LD, Kawachi I, et al. Depression and risk of sudden cardiac death and coronary heart disease in women: results from the Nurses’ Health Study. J Am Coll Cardiol. 2009;53:950–8.

    Article  PubMed  PubMed Central  Google Scholar 

  209. Buijsse B, Weikert C, Drogan D, Bergmann M, Boeing H. Chocolate consumption in relation to blood pressure and risk of cardiovascular disease in German adults. Eur Heart J. 2010;31:1616–23.

    Article  CAS  PubMed  Google Scholar 

  210. Djousse L, Hopkins PN, Arnett DK, Pankow JS, Borecki I, North KE, et al. Chocolate consumption is inversely associated with calcified atherosclerotic plaque in the coronary arteries: the NHLBI Family Heart Study. Clin Nutr. 2011;30:182–7.

    Article  PubMed  Google Scholar 

  211. Shonkoff JP, Garner AS; Committee on Psychosocial Aspects of Child and Family Health; Committee on Early Childhood, Adoption, and Dependent Care; Section on Developmental and Behavioral Pediatrics. The lifelong effects of early childhood adversity and toxic stress. Pediatrics. 2012;129:e232–e246.

  212. McCullough ML, Bostick RM, Mayo TL. Vitamin D gene pathway polymorphisms and risk of colorectal, breast, and prostate cancer. Annu Rev Nutr. 2009;29:111–32.

    Article  CAS  PubMed  Google Scholar 

  213. Bahadoran Z, Mirmiran P, Azizi F. Fast Food Pattern and CMD: A Review of Current Studies. Health Promotion Perspectives. 2015;5(4):231–40.

    Article  PubMed  Google Scholar 

  214. Mazidi M, Speakman J.R. Impact of obesity and ozone on the association between particulate air pollution and cardiovascular disease and stroke mortality among US adults. J Am Heart Assoc. 2018;7(11). https://doi.org/10.1161/JAHA.117.008006.

Download references

Acknowledgements

Not applicable

Funding

Not applicable

Author information

Authors and Affiliations

Authors

Contributions

ARE and AAF conceptualized the review study. ARE did the first draft. ARE and AAF reviewed the first draft. The authors read and approved the final manuscript.

Corresponding author

Correspondence to Ayodeji Ajayi.

Ethics declarations

Ethics approval and consent to participate

Not applicable

Consent for publication

All authors agree to the publication of this article

Competing interests

The authors declare that there are no competing interests

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Akhigbe, R., Ajayi, A. The impact of reactive oxygen species in the development of cardiometabolic disorders: a review. Lipids Health Dis 20, 23 (2021). https://doi.org/10.1186/s12944-021-01435-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12944-021-01435-7

Keywords