Crimson Publishers Publish With Us Reprints e-Books Video articles

Full Text

COJ Nursing & Healthcare

Evaluation of Enzymes Involved in the Production of Endogenous Reactive Oxygen Species and Role of Long-term Oxidative Stress in Pathobiology of Atherosclerosis with Common Oxidative Stress Markers

Leta Shiferaw Melaku*

Department of Biomedical Sciences,Ethiopia

*Corresponding author: Leta Shiferaw Melaku, Department of Biomedical Sciences, Ethiopia

Submission: March 27, 2020Published: January 28, 2021

DOI: 10.31031/COJNH.2021.07.000652

ISSN: 2577-2007
Volume7 Issue1

Abstract

Free radicals are molecules with an unpaired electron; thus, they are highly reactive. The human body deals with the pathological effects of free radicals by utilizing antioxidant system. The concentration and location of ROS are the main determinants of their effect. Although there are several sources of vascular free radicals, the enzyme NADPH oxidase is emerging as a strong candidate for the excessive ROS production. A multitude of studies provide evidence that an uncontrolled production of ROS is involved in the development and progression of cardiovascular disease like atherosclerosis. Atherosclerosis is characterized by the formation of intimal plaques. Monitoring and rapid detection of oxidative stress markers is necessary to combat the spread of various diseases.

Keywords:Free radicals; ROS; Antioxidant; Atherosclerosis; Oxidative stress markers

Introduction

As the key life-supporting element, oxygen was independently discovered by Priestly et al. [1] & Scheele et al. [2]. Within a few years of these seminal findings, oxygen toxic side effects that did not support life were also discovered [3]. The good and bad facets of oxygen are played out by its unique molecular structure [4]. The structural configuration of oxygen is a diradical can accept four electrons and the resultant one-step tetravalent reduction results in the formation of water, with a concurrent production of ATP. Ironically, if these four electrons are added one at a time, partially reduced forms of oxygen or free radicals are produced [5-7]. Free radicals can be defined as reactive chemical species having a single unpaired electron in an outer orbit [8]. This unstable configuration creates energy that can initiate autocatalytic reactions so that molecules to which they react are themselves converted into free radicals [9]. Although ROS (reactive oxygen species) more common in biological systems [9], free radicals also include RNS (reactive nitrogen species) [10]. The endogenous sources of ROS are the mainly by-products formed in the cells of aerobic organisms within mitochondria [11].
Furthermore, certain enzyme, neutrophils, eosinophil’s, macrophages, microsomes and peroxisomes are also sources of ROS [9,12-15]. It has been established that ROS can be both harmful and beneficial in biological systems depending on the environment [16,17]. At normal physiological levels, in phagocytic cells ROS plays a key role in cell-mediated immunity and microbiocidal activity [18,19]. In nonphagocytic cells, they are involved in a number of cellular signaling systems as well as in the induction or inhibition of cell proliferation [20-22]. In comparison, the rate of ROS production in nonphagocytic cells is only about one-third of that of phagocytic cells [23-26]. In contrast, at very high concentrations, ROS is often associated with the principle of oxidative stress [27]. The term oxidative stress is used to describe the condition of oxidative damage to a wide range of cellular structures as a result of an imbalance between free radical production and antioxidant defenses [28]. Short-term oxidative stress may occur in tissues injured by trauma, infection, heat injury, hypertoxia, toxins, and excessive exercise [29,30]. These harmful effects are balanced by the action of antioxidants [31]. However, in long-term oxidative stress, ROS have been implicated in the induction and complications of various cardiovascular diseases like atherosclerosis despite the presence of the cell’s antioxidant defense system [32,33].

Enzymes involved in the production of endogenous reactive oxygen species

Although the importance of ROS in vascular pathophysiology is quietly clear, recently there has been particular interest in the enzyme sources in the blood vessel wall [34]. This is due to the fact that enzymes are the most common sources of the production of endogenous ROS such as superoxide anion (O2), hydrogen peroxide (H2O2), and hydroxyl radicals (OH-) [35,36].

A. Nicotinamide-Adenine Dinucleotide Phosphate- Oxidase

Several enzymes that have potential to produce ROS are now recognized [37] and perhaps the most important one is NADPH oxidase, which consists of five subunits: p40PHOX, p47PHOX, p67PHOX, p22PHOX and Nox [38]. In resting cells, p47PHOX (47 kDa), p67PHOX (67 kDa) and p40PHOX (40 kDa) reside within the cytoplasm. On their stimulation, these polypeptides translocated to the inner face of the plasma membrane to form a fully active enzyme complex [39]. On the other hand, the plasma membrane contains two polypeptides, 22 kDa (p22PHOX) and 91 kDa (gp91-phox), which together make up flavocytochrome b558 [37]. This heterodimer contains a FAD group and two haeme groups, which enables the transfer of electrons from cytosolic NADPH across the membrane to molecular oxygen. Therefore, in the regulation of cytoplasmic NADPH oxidase activity, two G-proteins are associated [19,40,41]. Rap copurifies with gp91-phox, its exact role is still obscure, and p21rac, which is involved in the activation of NADPH oxidase. In the inactive state, p21rac is in association with a GDP-dissociation inhibition factor, Rho-GDI, but on stimulation dissociation takes place, and p21rac translocates to the plasma membrane, where it aids in the activation of the NADPH oxidase complex [42]. Moreover, NADPH oxidase activity is also regulated by phosphorylation of NADPH oxidase components. Phorbol myristate acetate (PMA) is commonly used to stimulate NADPH oxidase activity in cells and its action probably being mediated by protein kinase C [43,44]. It is likely that other kinases too are involved [39,45]. The modulation of intracellular calcium ions is also commonly used to activate the oxidase, and again a kinase may be involved in mediation of the calcium signal, although here direct stimulation has been suggested in some cell types [46]. Cardiovascular NADPH oxidase isoforms are also induced by hormones, hemodynamic forces, local metabolic changes and natural forces such as wall stress [47,48]. For instance, Angiotensin II increases NADPH-driven O2 production in cultured vascular smooth muscle cells and fibroblasts [49,50]. Thrombin, platelet-derived growth factor (PDGF), and TNF-α stimulate NADPH oxidase-dependent O2 production in vascular smooth muscle cells [51]. Interleukin-1, TNF-α, and PDGF increases NADPH-dependent O2 production in fibroblasts. Mechanical forces also stimulate NADPH oxidase activity in endothelial cells, and reoxygenation stimulates NADPH oxidase activity in cardiac myocytes [51-53].

B. Xanthine Oxidase/xanthine oxidoreductase

This molybdenum-and iron-containing flavoprotein catalyses the oxidation of hypoxanthine to xanthine and then to uric acid and molecular oxygen is the oxidant, whose products include O2 and H2O2 [46,54]. Unlike xanthine dehydrogenase, xanthine oxidases catalyse oxidation of uric acid xanthine in production of superoxide radicals [34]. In experimental animals with hypercholesterolemia, it is also capable of producing increased amounts of active radicals that directly leads to NO activity reduction [55]. Additional facts that support the role of xanthine oxidase in the process of atherogenesis are the following: 1) in patients with coronary syndrome the levels of this enzyme were found to be increased, the same applies to NAD(P)H; and 2) in young asymptomatic patients with familial hypercholesterolemia the increased activity of the enzyme is an early event [56]. It has been observed that in vessels of hypercholesterolemic patients, vasodilation is improved by the presence of allopurinol or oxypurinol, an inhibitor of the enzyme [34]. Therefore, this enzyme exists in plasma and endothelial cells but not in smooth muscle cells [34].

C. Myeloperoxidase

It is produced by activated phagocytes and uses H2O2 for the production of more powerful oxidative substances [41]. This enzyme, through NAD(P)H, leads to the production of HOCl and its analogs (substances related to endothelial injuries due to the action of H2O2) [57]. It is considered to participate in both cell-mediated immunity and microbiocidal activity [19,40] as well as in the process of atheromatosis, which is by the induction of oxidative modifications in low- and high-density lipoproteins [58]. This hypothesis is consistent with the results of clinical trials, according to which the levels of this enzyme and its products are elevated in patients with coronary syndrome. In contrast to human lesions, these oxidative products are absent in experimental animals with apolipoprotein E and LDL-receptor deficiency. The three mechanisms through which myeloperoxidase participate in oxidative modifications are NO consumption, LDL oxidation, and reaction with L-arginine for the production of NO synthase inhibitors. All of these are dependent on H2O2 [34]. Immunohistochemical studies have proved the presence of myeloperoxidase and HOCl in atherosclerotic lesions [59]. Therefore, both these substances participate in the modification of LDL and in atherogenesis.

D. Lipoxygenases

They are enzymes that catalyse the reaction of O2 with the polyunsaturated lipid acids, creating a family of biologically active lipids, such as prostaglandins, thromboxanes and leukotrienes, which participate in inflammatory reactions and increase the permeability of vessels [34]. In experimental models, 15-lipoxygenase induces LDL oxidation by enzymatic and non-enzymatic reactions [60]. Experimental animals with an absence of the 15-lipoxygenase gene or reduced expression of 5-lipoxygenase are protected from lesions like those found in animals with apolipoprotein E and LDL-receptor deficiency. Clinical data demonstrate that various genotypes of 5-lipoxygenase promoter are found in patients with atherosclerotic lesions or inflammation [61]. Whether lipoxygenases participate in atherogenesis through lipid oxidation or defensive modifications is under investigation.

The Role of Long-Term Oxidative Stress in Pathobiology of Atherosclerosis

Atherosclerosis is a multifactorial disease that involves the interplay of genetic and environmental factors and characterized by accumulation of cholesterol, infiltration of macrophages, proliferation of smooth muscle cells, and accumulation of connective tissue components and formation of thrombus [62,63]. It is the single largest cause of death and disability in the world [64] and most studies have shown that it starts early in life [65,66]. For instance, this disease appears earliest in the aorta (during fetal life), while it appears in the coronary arteries in the second decade and in the cerebral arteries in the third decade [65,66]. Furthermore, the growth of the lesion is abluminal in early stages of the disease, and the progress may vary from total cessation in some cases to very rapid with intervening periods of relative quiescence [67]. Hence, distinct clinical manifestations are seen depending on the type of vascular bed affected by atherosclerosis because it reduces the perfusion of a tissue [68,69]. Coronary lesions lead to myocardial ischemia or infarction [63]. Similarly, transient ischemic attacks and stroke are seen in the cerebral circulation, whereas intermittent claudication occurs in the peripheral circulation. Infarction of the gut produces lesions in the splanchnic circulation, while renal artery lesions result in ischemia due to reduced renal perfusion and damage the renal parenchyma, leading to uremia and eventually renal failure [70].
Several risk factors such as smoking, blood cholesterol, diabetes, physical inactivity and arterial hypertension are seen to contribute to the genesis of atherosclerosis; however, none of these factors are sufficient to produce an atherosclerotic lesion by themselves [71,72]. Evidence suggests that risk factors increase the risk of production of ROS from the endothelial cells, the smooth muscle cells and the adventitial cells of vasculature [73]. These ROS then oxidize cellular biomolecules to cause the atherosclerosis [74-76] as follows: the identification of the gaseous free radical as a major signal transducer molecule and EDRF [77] suggests oxy reduction reactions are important effector steps for autocrine or paracrine regulation of vessel tone, permeability, and structure in physiological or pathological conditions [78,79]. Therefore, the first physiological alteration in the pathobiology the problem is the impairment of the endothelium, which is manifested by enhanced vascular constriction and depressed dilatation of the vascular endothelium as well as excessive production of ROS [80]. The excess ROS generation, which is mainly due to NADPH oxidase activation, initiate vascular membrane lipid peroxidation that leads to inflammation and production of TNF-α via NF-κB induction [81,82] and other factors such as vascular adhesion molecule-1 (VCAM-1), monocyte chemoattractant protein-1 (MCP-1), endothelial-selectin, TGF-β1, Matrix metalloproteinase 9 (MMP9), iNOS and Mn-SOD [83-85].
Experimental evidence also supports crucial role for inflammatory reactions as a connection between risk factors for atherosclerotic disorder and pathophysiologic complexity of the disease [86]. TNF-α, which is one of inflammatory cytokines, is involved in commencement as well as development of atherosclerosis by inducing transcription factor nuclear factor- κB (NF-κB). In the process of atherosclerosis NF-κB induces the transcription of VCAM-1, ICAM-1, MCP-1, and E-selectin in smooth muscle/endothelial cells of the blood vessels [87]. TNF-α also depletes NO levels in the endothelium that leads to dysfunction of the endothelium [88,89]. In addition, TNF-α has been reported to cause apoptosis of the endothelial cells through dephosphorylation of protein kinase B (Akt) that leads feature endothelial damage [90,91]. Among the biomarkers of inflammation that is modulated by IL-6, IL-1 and TNF-α is C-reactive protein (CRP) [92] and evidence suggest that raised blood CRP level is an important predictor of CVD [93,94]. CRP is implicated in the advancement of atherosclerotic lesions by enhancing the production of VCAM-1, ICAM-1, selectins, and MCP-1 in the endothelium through induction of powerful constrictor of the vessels ET-1 and IL-6 [95,96].
Moreover, it ameliorates the synthesis of NO in the endothelium by depressing the transcription and translation of enzyme NO synthase [75]. It also plays a significant role in cooperating with the activities of other cytokines and factors. CRP induces the biochemical synthesis and physiological functions of PAI-1 in the endothelium [97]. PAI-1 is known to be actively involved in thrombosis during atherosclerosis process and inhibits destruction of the fibrin clot by suppressing plasminogen activation [98]. In atherosclerotic process, resistin, which exerts inflammatory reactions/vasoactive effects in the endothelium, also induces transcription of cellular factors such as VCAM-1 and MCP-1 [99]. Cells from endothelium exposed to resisting deplete the levels of TNF receptor-associated factor (TRAF-3) which is a well-known inhibitor of the endothelial activation [100]. It is also suggested that augmented resistin concentration causes a significant dysfunction of the endothelium through activation of endothelial system [101]. Furthermore, resistin exposure activates endothelial cells by increasing ET-1 release through induction of transcription of ET-1 indicating its role in the impairment of the endothelium [102]. Serum amyloid A (SAA) protein has also been implicated in the inflammatory reactions associated with atherosclerotic disorder and used as a biomarker for cardiac and vascular disorders as well as heart and vessels outcome [103].
The ROS also up-regulate atherosclerotic events such as cell infiltration, migration, adhesion and platelet activation [81,82,104] and facilitates the oxidation of low-density lipoprotein (LDL) and production of foam cells [105]. The transport of cholesterol regulated by ATP-binding cassette transporter A1 (ABCA1) and transport of oxidized LDL through CD36 regulate the excess of cholesterol ester in the macrophages, which result in formation of foam cells [105,106]. Apolipoprotein E as well as low-density lipoprotein (LDL)-receptor knock out animals display speedy atherosclerotic lesions [107,108] and they also have sizeable counts of macrophages/T cells in their plaques [106]. The elevated concentrations of factors involved in inflammatory pathway, namely TNF-α, MCP-1, Cox-2, TGF-β1, iNOS, and Mn-SOD in ApoE-deficient atherosclerotic mice [61,109], also proves the vascular inflammation as an integral process in the atherosclerotic pathophysiology [106]. Apart from excess foam cells, growth of smooth muscle/endothelial cells, collagens, matrix metalloproteinases, fibronectin, and elastin are also responsible for plaque development [75,76,80,110]. Leptin also increases the cellular growth as well as migration of cells of the endothelium [111] and cells of smooth muscle [112]. It also directly augments concentrations of monocyte colony-stimulating factor (MCSF) [113], increases cholesterol levels in hyperglycemia [114], and promotes new blood vessel formation [115]. It also induces the synthesis of MCP-1 in the cells of the aortic endothelium [116] and enhances the aggregation of the platelets and vascular thrombus formation through leptin receptor pathways [117,118]. Leptin also up regulates ET-1 as well as NO synthase biosynthesis in the endothelial cells and augments generation of free radicals and oxidants [117,118] that causes oxidative stress [119]. Depending on the histological picture, the lesions are classified into six types [120,121]. Type I contain atherogenic lipoproteins and infiltrates mononuclear leukocytes. The intima makes adaptive changes such as thickening.
This is seen in most people at birth. Type II has layers of macrophages or foam cells with SMC infiltration from the media into the intima. The gross lesion is designated as a fatty streak and is unique to the disease. Type III is an intermediary stage between types II and IV, with scattered coarse lipid granules or particles that disrupt the integrity of the SMC. Type IV lesions are characterized by typical atheromas containing a large extracellular lipid core and the abluminal growing atherosclerotic lesion. Type V lesions have atheromas with large extracellular lipid cores and the developing fibrous caps. There is an increase in the collagen and (more often) SMC content. Type V lesions are further classified into the Vb and Vc subtypes. Vb are characterized by largely calcified lesions, whereas the Type Vc contain more fibrous connective tissue, little lipid and no calcium [121]. Type VI lesions have ruptured atherosclerotic plaque with subsequent fissure formation or hematomas in the arterial lumen. As the thrombogenic lipid core comes into contact with the blood, thrombosis occurs due to platelet aggregation [63]. Inflammatory reactions are not only involved in progression of human vascular plaques generation but also have important role in the rupture of internal arterial plaques [106]. Generally, several factors are implicated in the rupture of internal arterial plaques including cytokines, cyclooxygenase-2, matrix metalloproteinases, and tissue factors [76,122,123].

Biologic Markers of Oxidative Modifications

Lipids, proteins, carbohydrates, and DNA are all susceptible to oxidative modifications of ROS [9,18]. Some modifications have direct functional effects, such as enzyme inhibition, with the remainder functionally silent indicators of increased ROS levels in the microenvironment [124]. The direct impact of the molecular modifications on the cell, organ and system’s ability to adapt to the elevated levels of ROS is an important contributor to the plausibility and validity of the marker, and its likelihood of emerging as a robust prognostic tool. However, it is challenged by the high reactivity and short half-life of many oxidative products as well as their variable specificity [125].

Lipids

Lipids are particularly susceptible targets of oxidation because of their abundant reactive double bonds [126]. Reactive oxygen species from the mitochondria, P450 enzymes, lipoxygenase and transition-metal catalysis are involved in lipid oxidation, or lipid peroxidation [127,128]. The ROS attack of the polyunsaturated fatty acids in the membrane and initiation of a self-propagating chain reaction results in altered fluidity and inactivation of critical membrane-bound receptors and enzymes [129]. Furthermore, the end-products of lipid peroxidation, such as the highly reactive secondary aldehyde products, isoketals from the isoprostane pathway, directly threaten the viability of tissues via their ability to covalently modify molecules that are critical to cell function [130]. Thus, lipid peroxidation is recognized as a crucial step in the pathogenesis of several CVD states including atherosclerosis [131,132].
The sensitivity of lipids to peroxidation, and its functional effects have made lipid peroxides good candidates as redox biomarkers [124]. The most frequently studied markers of lipid peroxidation are isoprostanes and malondialdehyde. However, the other markers include lipid hydroperoxides, fluorescent probes of lipid peroxidation and oxysterols [133,134]. Isoprostanes are prostaglandin-like substances that are produced independently of cyclooxygenase enzymes by ROS-induced peroxidation of arachidonic acid [135]. The most commonly measured members of the family are the F2-isoprostanes [134]. F2-isoprostanes are detectable in all biological fluids, reflecting baseline or ‘physiological’ levels of redox signalling [136]. They are substantially elevated in animal models of oxidant injury as well as human disease states characterized by elevated ROS [137]. They also increase in association with well-recognized risk factors such as cigarette smoking, hypercholesterolaemia, and diabetes mellitus [128]. Their causal role in human atherosclerosis is suggested by their effect to induce vasoconstriction [138], platelet aggregation [139], proliferation of VSMC [140] and their increased levels in atherosclerotic lesions [141].
Malondialdehyde (MDA), generated via peroxidation of polyunsaturated fatty acids, is also widely used to examine redox state [142]. Malondialdehyde-induced generation of lysinelysine cross-links in apolipoprotein B fractions of oxidized lowdensity lipoprotein (OxLDL) has been proposed to play a role in atherogenesis via impairing the action of macrophages [143]. Numerous studies have demonstrated the elevation of MDA in association with smoking and diabetes in both animals and humans. Malondialdehyde quantification therefore remains a useful biomarker in clinical research [124]. Another product of lipid peroxidation, 4-hydroxynonenal (4-HNE) appears to be particularly important for the regulation of vascular redox state in humans [144]. 4-hydroxynonenal is produced from the reaction of OH- with lipid structures, and is highly reactive with proteins, giving rise to a wide range of protein adducts [124]. Recent evidence suggests that 4-HNE produced in the vascular wall may exert paracrine effects on the neighbouring perivascular adipose tissue, leading to the activation of peroxisome proliferator-activated receptor-γ signalling in this fat depot [145]. As a result, perivascular fat releases the antioxidant adipokine, adiponectin, which exerts a paracrine effect back onto the vascular wall, reducing NADPH oxidase activity [145], and improving eNOS coupling. 4-Hydroxynonenal thus restores the balance between NO and O2 in the vascular endothelium [146].
This cascade also underlines the complexities of regulation of vascular redox state in humans, involving multiple intravascular feedback loops in addition to communication signals with other tissues, which host either pro- or antioxidant mechanisms depending on the underlying diseases state [124]. It also highlights that the oxidation products (used also as clinical biomarkers) may not always be ‘simple by-products’ of oxidation with no biological effects, but might play an active role in the regulation of vascular redox state, e.g., as rescue signals released from the vascular wall.

Protein

The direct, mostly reversible, functional effects of oxidative posttranslational modifications, like tyrosine nitration, protein carbonylation and S-glutathionylation, on many cellular proteins suggest proteins could be strong candidates for assessment of cellular redox haemostasis [147]. The nitration of protein tyrosines, which occurs through two predominant pathways, peroxynitrite and haeme peroxidase-dependent nitration [147,148] with steric effects, resulting in altered protein function, which is an important consequence of increased ROS. Many proteins including fibrinogen, plasmin, Apo A-I in the plasma, Apo B, Mn-superoxide dismutase (SOD) in the vessel wall, and creatine kinase (isoenzyme MM) as well as sarco/endoplasmic reticulum Ca2+-ATPase (SERCA) in the myocardium undergo nitration, with important functional effects [147].

Both free circulating 3-nitrotyrosine (3-NO2-Tyr), which possibly reflects the turnover of nitrated proteins with the modified amino acid not recycled for de novo protein synthesis, and total protein 3-NO2-Tyr, measured by hydrolyzing the protein fraction of the biological sample to its constituent amino acids, have been examined as biomarkers [149]. Although not used in clinical practice, the 3-NO2-Tyr has achieved a number of intermediate milestones, including demonstration of the levels as independent predictors of cardiovascular risk [150]. Protein carbonyls can be formed by the oxidation of a few amino acid side chains via the addition of aldehydes such as those generated from lipid peroxidation [151]. Carbonyl compounds are widely used markers of severe protein oxidation [152]. As a marker of oxidative damage to proteins, carbonyls have been shown to accumulate during aging, ischaemia/reperfusion [152], diabetes, and obesity [153].

Protein S-glutathionylation, the formation of a mixed disulphide bond between the reactive cysteine residue and the abundant glutathione is an excellent candidate for oxidative signalling due to its stability and reversibility (124). By conferring a 305 Da negatively charged adduct, it exerts steric effects on proteins similar to phosphorylation [154,155]. S-glutathionylation of critical cysteines plays a particularly important role in the cell membrane, mediating redox regulation of eNOS [156], the ryanodine receptor [157], SERCA [154], and the Na+-K+ pump [158], to name a few. In contrast to these, S-glutathionylation can also occur in non-critical cysteines without functional or regulatory effects [124]. Thus measuring ‘total S-glutathionylated proteins’ in serum, in a manner similar to that applied to protein nitrosylation, faces problems of both not representing S-glutathionylation at target tissues, as well as accounting for the subpopulation of ‘silently’ S-glutathionylated proteins. However, S-glutathionylation of the Na+-K+ pump in erythrocytes, which closely parallels in the myocardium in both animals and patients with heart failure [159], suggesting its biological validity as a circulatory marker in heart failure.

Advanced glycation end products

Advanced glycation end products are a class of molecules resulting from modifications of proteins or lipids that become nonenzymatically glycated and oxidized after contact with aldose sugars [124]. They form in vivo in hyperglycaemic environments and during the ageing process, and mediate vascular disease in diabetes [160]. Because of their severe instability, most of the advanced glycation end products are difficult to correctly analyse and are not practical for measurement as biomarkers in cardiovascular disease [124].

DNA

Reactive oxygen species can also mediate damage to all components of the DNA molecule, the purine and pyrimidine bases, as well as the deoxyribose backbone. Free radical induced damage to DNA in vivo can result in deleterious biological consequences such as the initiation and promotion of cancer [39]. One of the most abundant products of cellular DNA damage, 8-hydroxy-2′- deoxyguanosine (8-OHdG) can be detected by HPLC, and has been used as a redox biomarker, particularly in cancer research [124]. The levels of 8-OHdG have been found to be elevated in patients with CAD [161] and may also be useful for risk stratification in patients with subclinical cardiovascular disease, as shown for carotid atherosclerosis in a small study of haemodialysis patients [162].

Methodologies incorporating the technique of gas chromatography/mass spectrometry (GC/MS) have been also developed in recent years for measurement of free radical induced DNA damage. The use of GC/MS with selected-ion monitoring (SIM) facilitates unequivocal identification and quantitation of a large number of products of all four DNA bases produced in DNA by reactions with hydroxyl radical, hydrated electron, and H atom. Hydroxyl radical induced DNA-protein cross-links in mammalian chromatin, and products of the sugar moiety in DNA are also unequivocally identified [39]. The sensitivity and selectivity of the GC/MS-SIM technique enables the measurement of DNA base products even in isolated mammalian chromatin without the necessity of first isolating DNA, and despite the presence of histones. Recent studies revealed the usefulness of the GC/MS technique for chemical determination of free radical induced DNA damage in DNA as well as in mammalian chromatin under a vast variety of conditions of free radical production [163,164].

Conclusion

Free radicals play a dual role as both toxic and beneficial compounds, since they can be either harmful or helpful to the body. Many data support the notion that ROS released from nicotinamide adenine dinucleotide phosphate oxidase, myeloperoxidase, xanthine oxidase, lipoxygenase, nitric oxide synthase. When an overload of free radicals cannot gradually be destroyed, their accumulation in the body generates a phenomenon called oxidative stress. This process plays a major part in the development of various cardiovascular diseases such as atherosclerosis. ROS are key mediators of signaling pathways that underlie vascular inflammation in atherogenesis, starting from the initiation of fatty streak development, through lesion progression, to ultimate plaque rupture. Plaque rupture and thrombosis result in the acute clinical complications of myocardial infarction and stroke. Moreover, increased vascular production of ROS in atherosclerosis and common conditions predisposing to atherosclerosis such as hypercholesterolemia, hypertension, diabetes, and smoking likely contributes to development and progression of atherosclerosis by oxidative modification of LDL and by promoting endothelial dysfunction enhancing vascular inflammatory responses. Despite the biological plausibility of redox biomarkers as important adjuncts in diagnostic and prognostic armamentarium, their validation for clinical application has been slow and none have yet reached clinical use. It is likely that such pursuits will lead to a better understanding of these biological phenomena, and hopefully will provide new opportunities for therapeutic interventions.

References

  1. Priestly J (1775) Experiments and observations on different kinds of air. In: Simpkin (Ed.), The Discovery of Oxygen. Alembic Club Reprint No. 7. Marshall, Hamilton, Vol II, Sections III-V, London, UK, pp. 29-203.
  2. Scheele C (1777) Chemische abhandlung von der luft und dem Feuer, Upsala and Leipzig. In: Gurney Jackson (Ed.), The discovery of oxygen. Alembic Club Reprint No. 8, London, UK.
  3. Lavoisier A (1785) Alterations experienced by the air. Collection of Lavoisier's Memoirs. Read to the Society of Medicine. Reprinted as part of “Memories on Respiration and Transpiration of Animals” in ‘Masters of Scientific Thought. Gauthier-Villaus et cie, Paris, France.
  4. Cord J, Fridovich I (1969) Superoxide dismutase. An enzymatic function for erythrocuprein (hemecuprein). J Biol Chem 244(22): 6049-6055.
  5. Halliwell B (1987) Oxidants and human disease: some new concepts. FASEB J 1(5): 358-364.
  6. Singal P, Petkau A, Gerrard J, Hrushovetz S, Foerster J (1988) Free radicals in health and disease. Mol Cell Biochem 84(2):121-122.
  7. Kaul N, Iliskovic Ns, Hill M, Slezak J, Singal P (1993) Free radicals and the heart. J Pharmacol Toxicol Methods 30(2): 55-67.
  8. Riley P (1994) Free radicals in biology: oxidative stress and effects of ionizing radiation. Int J Rad Biol 65(1): 27-33.
  9. Rahman K (2007) Studies on free radicals, antioxidants, and co-factors. Clin Interv Aging 2(2): 219-236.
  10. Miller A, Budzyn K, Sobey C (2010) Vascular dysfunction in cerebrovascular disease: mechanisms and therapeutic intervention. Clin Sci 119(1): 1-17.
  11. Inoue M, Sato E, Nishikawa M, Park AM, Kira Y, et al. (2003) Mitochondrial generation of reactive oxygen species and its role in aerobic life. Curr Med Chem 10(23): 2495-2505.
  12. Valko M, Izakovic M, Mazur M, Rhodes C, Telser J (2004) Role of oxygen radicals in DNA damage and cancer incidence. Mol Cell Biochem 266(1-2): 37-56.
  13. Conner E, Grisham M (1996) Inflammation, free radicals, and antioxidants. Nutrition 12(4): 274-277.
  14. Valko M, Rhodes C, Moncol J, Izakovic M (2006) Free radicals, metals and antioxidants in oxidative stress-induced cancer. Chem Biol Interact 160(1): 1-40.
  15. Gupta M, Dobashi K, Greene E, Orak J, Singh I (1997) Studies on hepatic injury and antioxidant enzyme activities in rat sub-cellular organelles following in vivo ischemia and reperfusion. Mol Cell Biochem 176(1-2): 337-347.
  16. Lopaczynski W, Zeisel S (2001) Antioxidants, programmed cell death, and cancer. Nutr Res 21(1-2): 295-307.
  17. Glade M (2003) The role of reactive oxygen species in Health and Disease Northeast Regional Environmental Public Health Center University of Massachusetts. Amerst Nutrition 19: 401-403.
  18. Evans J, Goldfine I, Maddux B, Grodsky G (2003) Are oxidative stress-activated signaling pathways mediators of insulin resistance and beta-cell dysfunction? Diabetes 52(1): 1-8.
  19. Gaut J, Yeh G, Tran H, Byun J, Henderson J, et al. (2001) Neutrophils employ the myeloperoxidase system to generate antimicrobial brominating and chlorinating oxidants during sepsis. Proc Natl Acad Sci USA 98(21):11961-11966.
  20. Tyrrell R, Applegate L, Tromvoukis Y (1993) The proximal promoter region of the human heme oxygenase gene contains elements involved in stimulation of transcriptional activity by a variety of agents including oxidants. Carcinogenesis 14(4): 761-765.
  21. Um H, Orenstein J, Wahl S (1996) Fas mediates apoptosis in human monocytes by a reactive oxygen intermediate dependent pathway. J Immunol 156(9): 3469-3477.
  22. Bolwell G, Butt V, Davies D, Zimmerlin A (1995) The origin of the oxidative burst in plants. Free Radic Res 23(6): 517-523.
  23. Zweier J, Broderick R, Kuppusamy P, Gorman ST, Lutty G (1994) Determination of the mechanism of free radical generation in human aortic endothelial cells exposed to anoxia and reoxygenation. J Biol Chem 269(39): 24156-24162.
  24. Thannickal V, Fanburg B (1995) Activation of an H2O2- generating NADH oxidase in human lung fibroblasts by transforming growth factor beta 1. J Biol Chem 270(51): 30334-30338.
  25. Bae Y, Kang S, Seo M, Baines I, Tekle E, et al. (1997) Epidermal growth factor (EGF)-induced generation of hydrogen peroxide. Role in EGF receptor-mediated tyrosine phosphorylation. J Biol Chem 272(1): 217-221.
  26. Suh Y, Arnold R, Lassegue B, Shi J, Xu X, et al. (1999) Cell transformation by the superoxide-generating oxidase Mox1. Nature 401(6748): 79-82.
  27. Cross C, Halliwell B, Borish E, Pryor W, Ames B, et al. (1987) Oxygen radicals and human disease. Ann intern med 107(4): 526-545.
  28. Rock C, Jacob R, Bowen P (1996) Update on the biological characteristics of the antioxidant micronutrients Vitamin C, Vitamin E and the carotenoids. J Am Diet Assoc 96(7): 693-702.
  29. Cord J (2000) The evolution of free radicals and oxidative stress. Am J Med 108(8): 652-659.
  30. Lobo V, Patil A, Phatak A, Chandra N (2010) Free radicals, antioxidants and functional foods: Impact on human health. Pharmacogn Rev 4(8):118-126.
  31. Halliwell B (1996) Antioxidants in human health and disease. Ann Rev Nutr 16: 33-50.
  32. Rao A, Bharani M, Pallavi V (2006) Role of antioxidants and free radicals in health and disease. Adv Pharmacol Toxicol 7: 29-38.
  33. Rahman K (2003) Garlic and aging: new insights into an old remedy. Ageing Res Rev 2(1): 39-56.
  34. Vogiatzi G, Stousoulis D, Stefanadis C (2009) The role of oxidative stress in atherosclerosis. Hellenic J Cardiol 50(5): 402-409.
  35. Gutteridge J (1995) Lipid peroxidation and antioxidants as biomarkers of tissue damage. Clin Chem 41(12 Pt 2): 1819-1828.
  36. Cadenas E, Sies H (1998) The lag phase. Free Rad Res 28(6): 601-609.
  37. Babior B (1999) NADPH oxidase: an update. Blood 93(5): 1464-1476.
  38. Guzik T, West N, Black E, McDonald D, Ratnatunga C, et al. (2000) Vascular superoxide production by NAD(P)H oxidase. Association with endothelial dysfunction and clinical risk factors. Circ Res 86(9): E85-90.
  39. Genestra M (2007) Oxyl radicals, redox-sensitive signalling cascades and antioxidants. Cell Signal 19(9): 1807-1819.
  40. Klebanoff S, Rosen H (1978) The role of myeloperoxidase in the microbicidal activity of polymorphonuclear leukocytes. Ciba Found Symp 65: 263-284.
  41. Gaut J, Byun J, Tran H, Lauber W, Carroll J, et al. (2002) Myeloperoxidase produces nitrating oxidants in vivo. J Clin Invest 109(10): 1311-1319.
  42. Henderson L, Chappell J, Jones O (1988) Internal pH changes associated with activity of NADPH oxidase of human neutrophils. Further evidence for the presence of an H+ conducting channel. Biochem J 251(2): 563-567.
  43. Morgan D, Cherny V, Finnegan A, Bollinger J, Gelb M, et al. (2007) Sustained activation of proton channels and NADPH oxidase in human eosinophils and murine granulocytes requires PKC but not cPLA2 alpha activity. J Physiol 579(Pt 2): 327-344.
  44. Ezeamuzie C, Taslim N (2006) Reactive oxygen species mediate phorbol ester-stimulated cAMP response in human eosinophils. Eur J Pharmacol 543(1-3): 174-180.
  45. Roder J, Helfand S, Werkmeister J, McGarry R, Beaumont T, et al. (1982) Oxygen intermediates are triggered early in the cytolytic pathway of human NK cells. Nature 298(5874): 569-572.
  46. McNally J, Saxena A, Cai H, Dikalov S, Harrison D (2005) Regulation of xanthine oxidoreductase protein expression by hydrogen peroxide and calcium. Arterioscler Thromb Vasc Biol 25(8): 1623-1628.
  47. Rpshchupkin D, Berzhitskaia V, Murina M (1998) Difference in inhibitory actions of products of the myeloperoxidase- catalyzed reaction on initial aggregation of activated platelets. Biofizika 43(2): 323-328.
  48. Griendling K, Sorescu D, Lassegue B, Fukai MU (2000) Modulation of protein kinase activity and gene expression by reactive oxygen species and their role in vascular physiology and pathophysiology. Arterioscler Thromb Vasc Biol 20(10): 2175-2183.
  49. Adachi T, Pimentel D, Heibeck T, Hou X, Lee Y, et al. (2004) S-glutathiolation of Ras mediates redox-sensitive signaling by angiotensin II in vascular smooth muscle cells. J Biol Chem 279(28): 29857-29862.
  50. Hanna I, Taniyama Y, Szocs K, Rocic P, Griendling K (2002) NAD(P)H oxidase-derived reactive oxygen species as mediators of angiotensin II signaling. Antioxid Redox Signal 4(6): 899-914.
  51. Fortuno A, Jose G, Moreno M, Diez J, Zalba G (2005) Oxidative stress and vascular remodelling. Exp Physiol 90(4): 457-462.
  52. Devillard L, Vandroux D, Tissier C, Brochot A, Voisin S, et al. (2006) Tubulin ligands suggest a microtubule-NADPH oxidase relationship in postischemic cardiomyocytes. Eur J Pharmacol 548(1-3): 64-73.
  53. Ateghang B, Wartenberg M, Gassmann M, Sauer H (2006) Regulation of cardiotrophin-1 expression in mouse embryonic stem cells by HIF-1alpha and intracellular reactive oxygen species. J Cell Sci 119(Pt 6): 1043-1052.
  54. Jones O, Hancock J (2000) Free radicals in inflammation. In: Winyard PG, Blake DR, Evans CH (Eds.), Birkha, Switzerland, pp. 21-46.
  55. Antoniades C, Tousoulis D, Marinou K, Stefanadi E, Ntarladimas I, et al. (2006) Effects of lipid profile on forearm hyperemic response in young subjects. Hellenic J Cardiol 47(3): 152-157.
  56. Spiekermann S, Landmesser U, Dikalov S, Bredt M, Gamez G, et al. (2003) Electron spin resonance characterization of vascular xanthine, oxidase activity in patients with coronary artery disease NAD(P)H. relation to endothelium dependent vasolidation. Circulation 107(10): 1383-1389.
  57. Bergt C, Pennathur S, Fu X, Byun J, Brien K, et al. (2004) The myeloperoxidase product hypochlorous acid oxidizes HDL in the human artery wall and impairs ABCA-1 dependent cholesterol transport. Proc Natl Acad Sci USA 101(35): 13032-13037.
  58. Daugherty A, Dunn J, Rateri D, Heinecke JW (1994) Myeloperoxidase, a catalyst for lipoprotein oxidation, is expressed in human atherosclerotic lesions. J Clin Invest 94(1): 437-444.
  59. Pennathur S, Bergt C, Shao B, Byun J, Kassim S, et al. (2004) Human atherosclerotic intima and blood of patients with established coronary artery disease contain HDL damaged by reactive nitrogen species. J Biol Chem 279(41): 42977-42983.
  60. Tousoulis D, Böger R, Antoniades C, Siasos G, Stefanadi E, et al. (2007) Mechanisms of disease: L-arginine in coronary atherosclerosis-a clinical perspective. Nat Clin Pract Cardiovasc Med 4(5): 274-283.
  61. Husain K, Suarez E, Isidro A, Ferder L (2010) Effects of paricalcitol and enalapril on atherosclerotic injury in mouse aortas. Am J Nephrol 32(4): 296-304.
  62. Turunen M, Hiltunen M, Herttuala S (1999) Gene therapy for angiogenesis, restenosis and related diseases. Exp Gerontol 34(4): 567-574.
  63. Singh R, Mengi S, Xu Y, Arneja A, Dhalla N (2002) Pathogenesis of atherosclerosis: A multifactorial process. Exp Clin Cardiol 7(1): 40-53.
  64. Hegele R (1997) The genetic basis of atherosclerosis. Int J Clin Lab Res 27(1): 2-13.
  65. Palinski W, Napoli C (1999) Pathophysiological events during pregnancy influence the development of atherosclerosis in humans. Trends Cardiovasc Med 9(7): 205-214.
  66. Napoli C, Armiento F, Corso G, Ambrosio G, Palumbo G, et al. (1997) Occurrence of the same peroxidative compounds in low density lipoprotein and in atherosclerotic lesions from a homozygous familial hypercholesterolemic patient: a case report. Int J Cardiol 62(1): 77-85.
  67. Libby P, Schoenbeck U, Mach F, Selwyn A, Ganz P (1998) Current concepts in cardiovascular pathology: the role of LDL cholesterol in plaque rupture and stabilization. Am J Med 104(2A): 14S-18S.
  68. Smedby O, Johansson J, Molgaard J, Olsson A, Walldius A, et al. (1995) Predilection of atherosclerosis for the inner curvature in the femoral artery. A digitized angiography study. Arterioscler Thromb Vasc Biol 15(7): 912-917.
  69. Smedby O (1996) Geometric risk factors for atherosclerosis in the aortic bifurcation: a digitized angiography study. Ann Biomed Eng 24(4): 481-488.
  70. Davies M, Fulton G, Hagen P (1995) Clinical biology of nitric oxide. Br J Surg 82(12): 1598-1610.
  71. Novo S, Failla G, Liquori M, Longo B, Gennaro C, et al. (1991) Vascular damage in arterial hypertension: its noninvasive assessment. Cardiologia 36(12 Suppl 1): 323-337.
  72. Novo S, Avellone G, Garbo V, Abrignani M, Liquori M, et al. (1992) Prevalence of risk factors in patients with peripheral arterial disease. A clinical and epidemiological evaluation. Int Angiol 11(3): 218-229.
  73. Gozin A, Franzini E, Andrieu V, Costa L, Labelle E, et al. (1998) Reactive oxygen species activate focal adhesion kinase, paxillin and p130cas tyrosine phosphorylation in endothelial cells. Free Radic Biol Med 25(9): 1021-1032.
  74. Hamza S, Dyck J (2014) Systemic and renal oxidative stress in the pathogenesis of hypertension: modulation of long-term control of arterial blood pressure by resveratrol. Front Physiol 5: 292.
  75. Verma S, Wang C, Li S, Dumont A, Fedak P, et al. (2002) A self-fulfilling prophecy: C-reactive protein attenuates nitric oxide production and inhibits angiogenesis. Circulation 106(8): 913-919.
  76. Libby P (2002) Inflammation in atherosclerosis. Nature 420(6917): 868-874.
  77. Moncada S, Palmer R, Higgs E (1991) Nitric oxide: physiology, pathophysiology and pharmacology. Pharmacol Rev 43(2): 109-142.
  78. Gryglewsky R, Palmer R, Moncada S (1986) Superoxide anion is involved in the breakdown of endothelium-derived vascular relaxing factor. Nature 320: 454-456.
  79. Rubanyi G, Vanhoutte P (1986) Superoxide anions and hyperoxia inactivate endothelium-derived relaxing factor. Am J Physiol 250(5 Pt 2): H822-H827.
  80. Ross R (1999) Atherosclerosis-an inflammatory disease. N Engl J Med 340(2): 115-126.
  81. Ungvari Z, Csiszar A, Kaminski P, Wolin M, Koller A (2004) Chronic high pressure-induced arterial oxidative stress: involvement of protein kinase C-dependent NAD(P)H oxidase and local reninangiotensin system. Am J Pathol 165(1): 219-226.
  82. Zhang L, Ma Y, Zhang J, Cheng J, Du J (2005) A new cellular signaling mechanism for angiotensin II activation of NF-kappaB: An Ikappa B independent, RSK-mediated phosphorylation of p65. Arterioscler Thromb Vasc Biol 25(6): 1148-1153.
  83. Jiang X, Zeng H, Guo Y, Zhou Z, Tang B, et al. (2004) The expression of matrix metalloproteinases-9, transforming growth factor-beta1 and transforming growth factor-beta receptor I in human atherosclerotic plaque and their relationship with plaque stability. Chin Med J (Engl) 117(12): 1825-1829.
  84. Gustafsson S, Lind L, Söderberg S, Zilmer M, Hulthe J, et al. (2013) Oxidative stress and inflammatory markers in relation to circulating levels of adiponectin. Obesity (Silver Spring) 21(7): 1467-1473.
  85. Lu W, Jiang J, Hu J, Wang J, Zheng M (2015) Curcumin protects against lipopolysaccharide-induced vasoconstriction dysfunction via inhibition of thrombospondin-1 and transforming growth factor-β Exp Ther Med 9(2): 377-383.
  86. Libby P, Okamoto Y, Rocha V, Folco E (2010) Inflammation in atherosclerosis: transition from theory to practice. Circ J 74(2): 213-220.
  87. Ouchi N, Kihara S, Arita Y, Maeda K, Kuriyama H, et al. (1999) Novel modulator for endothelial adhesion molecules: adipocyte-derived plasma protein adiponectin. Circulation 100(25): 2473-2476.
  88. Bhagat K, Vallance P (1997) Inflammatory cytokines impair endothelium-dependent dilatation in human veins in vivo. Circulation 96(9): 3042-3047.
  89. Wang P, Ba Z, Chaudry I (1994) Administration of tumor necrosis factor-alpha in vivo depresses endothelium-dependent relaxation. Am J Physiol 266(6 Pt 2): H2535-H2541.
  90. Choy J, Granville D, Hunt D, Manus B (2001) Endothelial cell apoptosis: biochemical characteristics and potential implications for atherosclerosis. J Mol Cell Cardiol 33(9): 1673-1690.
  91. Hermann C, Assmus B, Urbich C, Zeiher A, Dimmeler S (2000) Insulin-mediated stimulation of protein kinase Akt: A potent survival signaling cascade for endothelial cells. Arterioscler Thromb Vasc Biol 20(2): 402-409.
  92. Yudkin J, Stehouwer C, Emeis J, Coppack S (1999) C-reactive protein in healthy subjects: associations with obesity, insulin resistance, and endothelial dysfunction: a potential role for cytokines originating from adipose tissue? Arterioscler Thromb Vasc Biol 19(4): 972-978.
  93. Ridker P, Buring J, Cook N, Rifai N (2003) C-reactive protein, the metabolic syndrome, and risk of incident cardiovascular events: an 8-year follow-up of 14 719 initially healthy American women. Circulation 107: 391-397.
  94. Visser M, Bouter L, McQuillan G, Wener M, Harris T (1999) Elevated C-reactive protein levels in overweight and obese adults. JAMA 282(22): 2131-2135.
  95. Pasceri V, Wu H, Willerson J, Yeh E (2000) Modulation of vascular inflammation in vitro and in vivo by peroxisome proliferator activated receptor-gamma activators. Circulation 101(3): 235-238.
  96. Pepys M, Hirschfield G (2003) C-reactive protein: a critical update. J Clin Invest 111(12): 1805-1812.
  97. Devaraj S, Xu D, Jialal I (2003) C-reactive protein increases plasminogen activator inhibitor-1 expression and activity in human aortic endothelial cells: implications for the metabolic syndrome and atherothrombosis. Circulation 107(3): 398-404.
  98. Kohler H, Grant P (2000) Plasminogen-activator inhibitor type 1 and coronary artery disease. N Engl J Med 342(24): 1792-1801.
  99. Steppan C, Brown E, Wright C, Bhat S, Banerjee R, et al. (2001) A family of tissue-specific resistin-like molecules. Proc Natl Acad Sci USA 98(2): 502-506.
  100. Calabro P, Samudio I, Willerson J, Yeh E (2004) Resistin promotes smooth muscle cell proliferation through activation of extracellular signal-regulated kinase 1/2 and phosphatidylinositol 3-kinase pathways. Circulation 110(21): 3335-3340.
  101. Verma S, Li S, Wang C, Fedak P, Li R, et al. (2003) Resistin promotes endothelial cell activation: further evidence of adipokine-endothelial interaction. Circulation 108(6): 736-740.
  102. Lau D, Dhillon B, Yan H, Szmitko P, Verma S (2005) Adipokines: molecular links between obesity and atheroslcerosis. Am J Physiol Heart Circ Physiol 288(5): H2031-H2041.
  103. Johnson B, Kip K, Marroquin O, Ridker P, Kelsey S, et al. (2004) Serum amyloid A as a predictor of coronary artery disease and cardiovascular outcome in women: The National Heart, Lung, and Blood Institute-Sponsored Women’s Ischemia Syndrome Evaluation (WISE). Circulation 109(6): 726-732.
  104. Brasier A (2010) The nuclear factor-kappaB-interleukin-6 signalling pathway mediating vascular inflammation. Cardiovasc Res 86(2): 211-218.
  105. Allahverdian S, Pannu P, Francis G (2012) Contribution of monocyte derived macrophages and smooth muscle cells to arterial foam cell formation. Cardiovasc Res 95(2): 165-172.
  106. Husain K, Hernandez W, Ansari R, Ferder L (2015) Inflammation, oxidative stress and renin angiotensin system in atherosclerosis. World J Biol Chem 6(3): 209-217.
  107. Yang H, Roberts L, Shi M, Zhou L, Ballard B, et al. (2004) Retardation of atherosclerosis by overexpression of catalase or both Cu/Zn-superoxide dismutase and catalase in mice lacking apolipoprotein E. Circ Res 95(11): 1075-1081.
  108. Zhang S, Reddick R, Piedrahita J, Maeda N (1992) Spontaneous hypercholesterolemia and arterial lesions in mice lacking apolipoprotein E. Science 258(5081): 468-471.
  109. Martinez ES, Husain K, Ferder L (2014) Adiponectin expression and the cardioprotective role of the vitamin D receptor activator paricalcitol and the angiotensin converting enzyme inhibitor enalapril in ApoE-deficient mice. Ther Adv Cardiovasc Dis 8(6): 224-236.
  110. Obikane H, Abiko Y, Ueno H, Kusumi Y, Esumi M, et al. (2010) Effect of endothelial cell proliferation on atherogenesis: a role of p21(Sdi/Cip/Waf1) in monocyte adhesion to endothelial cells. Atherosclerosis 212(1): 116-122.
  111. Park H, Kwon H, Lim H, Hong B, Lee J, et al. (2001) Potential role of leptin in angiogenesis: leptin induces endothelial cell proliferation and expression of matrix metalloproteinases in vivo and in vitro. Exp Mol Med 33(2): 95-102.
  112. Artwohl M, Roden M, Hölzenbein T, Freudenthaler A, Waldhäusl W, et al. (2002) Modulation by leptin of proliferation and apoptosis in vascular endothelial cells. Int J Obes Relat Metab Disord 26(4): 577-580.
  113. Loffreda S, Yang S, Lin H, Karp C, Brengman M, et al. (1998) Leptin regulates proinflammatory immune responses. FASEB J 12(1): 57-65.
  114. Rourke LO, Gronning L, Yeaman S, Shepherd P (2002) Glucose-dependent regulation of cholesterol ester metabolism in macrophages by insulin and leptin. J Biol Chem 277(45): 42557-42562.
  115. Honigmann MS, Nath A, Murakami C, Cardeña G, Papapetropoulos A, et al. (1998) Biological action of leptin as an angiogenic factor. Science 281(5383): 1683-1686.
  116. Yamagishi S, Edelstein D, Du X, Kaneda Y, Guzmán M, et al. (2001) Leptin induces mitochondrial superoxide production and monocyte chemoattractant protein-1 expression in aortic endothelial cells by increasing fatty acid oxidation via protein kinase A. J Biol Chem 276(27): 25096-25100.
  117. Konstantinides S, Schafer K, Loskutoff D (2001) The prothrombotic effects of leptin possible implications for the risk of cardiovascular disease in obesity. Ann N Y Acad Sci 947: 134-141.
  118. Cooke J, Oka R (2002) Does leptin cause vascular disease? Circulation 106(15): 1904-1905.
  119. Bouloumie A, Marumo T, Lafontan M, Busse R (1999) Leptin induces oxidative stress in human endothelial cells. FASEB J 13(10): 1231-1238.
  120. Stary H, Chandler A, Glagov S, Guyton J, Insull W, et al. (1994) A definition of initial, fatty streak, and intermediate lesions of atherosclerosis. A report from the Committee on Vascular Lesions of the Council on Arteriosclerosis, American Heart Association. Circulation 89(5): 2462-2478.
  121. Stary H, Chandler A, Dinsmore R, Fuster V, Glagov S, et al. (1995) A definition of advanced types of atherosclerotic lesions and a histological classification of atherosclerosis. A report from the Committee on Vascular Lesions of the Council on Arteriosclerosis, American Heart Association. Circulation 92(5): 1355-1374.
  122. Scott J (2002) The pathogenesis of atherosclerosis and new opportunities for treatment and prevention. J Neural Transm Suppl (63): 1-17.
  123. Stoll G, Bendszus M (2006) Inflammation and atherosclerosis: novel insights into plaque formation and destabilization. Stroke 37(7): 1923-1932.
  124. Galougahi K, Antoniades C, Nicholls S, Channon K, Figtree G (2015) Redox biomarkers in cardiovascular medicine. European Heart Journal 36(25): 1576-1582.
  125. Lee R, Margaritis M, Channon K, Antoniades C (2012) Evaluating oxidative stress in human cardiovascular disease: methodological aspects and considerations. Curr Med Chem 19(16): 2504-2520.
  126. Porter N, Caldwell S, Mills K (1995) Mechanisms of free radical oxidation of unsaturated lipids. Lipids 30: 277-290.
  127. Smith K, Shepherd J, Wakil A, Kilpatrick E (2011) A comparison of methods for the measurement of 8-isoPGF(2alpha): a marker of oxidative stress. Ann Clin Biochem 48(Pt 2): 147-154.
  128. Morrow J (2005) Quantification of isoprostanes as indices of oxidant stress and the risk of atherosclerosis in humans. Arterioscler Thromb Vasc Biol 25(2): 279-286.
  129. Mylonas C, Kouretas D (1999) Lipid peroxidation and tissue damage. In Vivo 13(3): 295-309.
  130. Gutteridge J, Halliwell B (1990) The measurement and mechanism of lipid peroxidation in biological systems. Trends Biochem Sci 15(4): 129-135.
  131. Steinberg D (1997) Low density lipoprotein oxidation and its pathobiological significance. J Biol Chem 272(34): 20963-20966.
  132. Chisolm G, Steinberg D (2000) The oxidative modification hypothesis of atherogenesis: an overview. Free Radic Biol Med 28(12): 1815-1826.
  133. Ho E, Galougahi K, Liu C, Bhindi R, Figtree G (2013) Biological markers of oxidative stress: applications to cardiovascular research and practice. Redox Biol 1(1): 483-491.
  134. Montuschi P, Barnes P, Roberts L (2004) Isoprostanes: markers and mediators of oxidative stress. FASEB J 18(15): 1791-1800.
  135. Morrow J, Awad J, Boss H, Blair I, Roberts L (1992) Non-cyclooxygenase-derived prostanoids (F2-isoprostanes) are formed in situ on phospholipids. Proc Natl Acad Sci USA 89(22): 10721-10725.
  136. Wu T, Rifai N, Roberts L, Willett W, Rimm E (2004) Stability of measurements of biomarkers of oxidative stress in blood over 36 hours. Cancer Epidemiol Biomarkers Prev 13(8): 1399-1402.
  137. Griffiths H, Moller L, Bartosz G, Bast A, Freddari CB, et al. (2002) Biomarkers. Mol Aspects Med 23(1-3): 101-208.
  138. Kromer B, Tippins J (1996) Coronary artery constriction by the isoprostane 8-epi prostaglandin F2 alpha. Br J Pharmacol 119(6): 1276-1280.
  139. Patrono C, FitzGerald G (1997) Isoprostanes: potential markers of oxidant stress in atherothrombotic disease. Arterioscler Thromb Vasc Biol 17(11): 2309-2315.
  140. Takahashi K, Nammour T, Fukunaga M, Ebert J, Morrow J, et al. (1992) Glomerular actions of a free radical-generated novel prostaglandin, 8-epi-prostaglandin F2 alpha, in the rat. Evidence for interaction with thromboxane A2 receptors. J Clin Invest 90(1): 136-141.
  141. Gniwotta C, Morrow J, Roberts L, Kuhn H (1997) Prostaglandin F2-like compounds, F2-isoprostanes, are present in increased amounts in human atherosclerotic lesions. Arterioscler Thromb Vasc Biol 17(11): 3236-3241.
  142. Mallat Z, Philip I, Lebret M, Chatel D, Maclouf J, et al. (1998) Elevated levels of 8-iso-prostaglandin F2alpha in pericardial fluid of patients with heart failure: a potential role for in vivo oxidant stress in ventricular dilatation and progression to heart failure. Circulation 97(16): 1536-1539.
  143. Uchida K (2000) Role of reactive aldehyde in cardiovascular diseases. Free Radic Biol Med 28(12): 1685-1696.
  144. Walter M, Jacob R, Jeffers B, Ghadanfar M, Preston G, et al. (2004) Serum levels of thiobarbituric acid reactive substances predict cardiovascular events in patients with stable coronary artery disease: a longitudinal analysis of the PREVENT study. J Am Coll Cardiol 44(10): 1996-2002.
  145. Antonopoulos A, Margaritis M, Coutinho P, Shirodaria C, Psarros C, et al. (2014) Adiponectin as a link between type 2 diabetes mellitus and vascular NADPH-oxidase activity in the human arterial wall: the regulatory role of perivascular adipose tissue. Diabetes 64(6): 2207-2219.
  146. Margaritis M, Antonopoulos A, Digby J, Lee R, Reilly S, et al. (2013) Interactions between vascular wall and perivascular adipose tissue reveal novel roles for adiponectin in the regulation of endothelial nitric oxide synthase function in human vessels. Circulation 127(22): 2209-2221.
  147. Peluffo G, Radi R (2007) Biochemistry of protein tyrosine nitration in cardiovascular pathology. Cardiovasc Res 75(2): 291-302.
  148. Schopfer F, Baker P, Freeman B (2003) NO-dependent protein nitration: a cell signalling event or an oxidative inflammatory response? Trends Biochem Sci 28(12): 646-654.
  149. Duncan M (2003) A review of approaches to the analysis of 3-nitrotyrosine. Amino Acids 25(3-4): 351-361.
  150. Shishehbor M, Aviles R, Brennan M, Fu X, Goormastic M, et al. (2003) Association of nitrotyrosine levels with cardiovascular disease and modulation by statin therapy. JAMA 289(13): 1675-1680.
  151. Grimsrud P, Xie H, Griffin T, Bernlohr D (2008) Oxidative stress and covalent modification of protein with bioactive aldehydes. J Biol Chem 283(32): 21837-21841.
  152. Donne I, Rossi R, Colombo R, Giustarini D, Milzani A (2006) Biomarkers of oxidative damage in human disease. Clin Chem 52(4): 601-623.
  153. Bollineni R, Fedorova M, Bluher M, Hoffmann R (2014) Carbonylated plasma proteins as potential biomarkers of obesity induced type 2 diabetes mellitus. J Proteome Res 13(11): 5081-5093.
  154. Adachi T, Weisbrod R, Pimentel D, Ying J, Sharov V, et al. (2004) S-Glutathiolation by peroxynitrite activates SERCA during arterial relaxation by nitric oxide. Nat Med 10(11): 1200-1207.
  155. Figtree G, Keyvan Karimi G, Liu C, Rasmussen H (2012) Oxidative regulation of the Na(+)-K(+) pump in the cardiovascular system. Free Radic Biol Med 53(12): 2263-2268.
  156. Chen C, Wang T, Varadharaj S, Reyes L, Hemann C, et al. (2010) S-Glutathionylation uncouples eNOS and regulates its cellular and vascular function. Nature 468(7327): 1115-1118.
  157. Aracena PP, Goonasekera S, Gilman C, Dirksen R, Hidalgo C, et al. (2006) Identification of cysteines involved in S-nitrosylation, S-glutathionylation, and oxidation to disulfides in ryanodine receptor type 1. J Biol Chem 281(52): 40354-40368.
  158. Figtree G, Liu C, Bibert S, Hamilton E, Garcia A, et al. (2009) Reversible oxidative modification: a key mechanism of Na+-K+ pump regulation. Circ Res 105(2): 185-193.
  159. Liu C, Fry N, Karimi Galougahi K, Rasmussen H, et al. (2012) Glutathionylation of erythrocyte Na-K Pump in heart failure: a novel biomarker that reflects a key oxidative abnormality in the heart. Circulation 126(Supp_21): A12793.
  160. Goldin A, Beckman J, Schmidt A, Creager M (2006) Advanced glycation end products: sparking the development of diabetic vascular injury. Circulation 114(6): 597-605.
  161. Kaya Y, Cebi A, Soylemez N, Demir H, Alp H, et al. (2012) Correlations between oxidative DNA damage, oxidative stress and coenzyme Q10 in patients with coronary artery disease. Int J Med Sci 9(8): 621-626.
  162. Ari E, Kaya Y, Demir H, Cebi A, Alp H, et al. (2011) Oxidative DNA damage correlates with carotid artery atherosclerosisin hemodialysis patients. Hemodial Int 15(4): 453-459.
  163. Jaruga P, Rodriguez H, Dizdaroglu M (2001) Measurement of 8-hydroxy-2’-deoxyadenosine in DNA by liquid chromatography/mass spectrometry. Free Radic Biol Med 31(3): 336-344.
  164. Dizdaroglu M, Jaruga P, Birincioglu M, Rodriguez H (2002) Free radical-induced damage to DNA: mechanisms and measurement. Free Radic Biol Med 32(11): 1102-1115.

© 2021 Leta Shiferaw Melaku. This is an open access article distributed under the terms of the Creative Commons Attribution License , which permits unrestricted use, distribution, and build upon your work non-commercially.