当前位置: 首页 > 期刊 > 《动脉硬化血栓血管生物学》 > 2005年第1期 > 正文
编号:11275477
Oxidative Stress and Vascular Disease
http://www.100md.com 《动脉硬化血栓血管生物学》
     From the Carolina Cardiovascular Biology Center, Department of Medicine, University of North Carolina, Chapel Hill.

    Correspondence to Marschall S. Runge, MD, PhD, Department of Medicine, 3033 Old Clinic Bldg., University of North Carolina, Chapel Hill, NC 27599-7005. E-mail mrunge@med.unc.edu

    Abstract

    Growing evidence indicates that chronic and acute overproduction of reactive oxygen species (ROS) under pathophysiologic conditions is integral in the development of cardiovascular diseases (CVD). These ROS can be released from nicotinamide adenine dinucleotide (phosphate) oxidase, xanthine oxidase, lipoxygenase, mitochondria, or the uncoupling of nitric oxide synthase in vascular cells. ROS mediate various signaling pathways that underlie vascular inflammation in atherogenesis: from the initiation of fatty streak development through lesion progress to ultimate plaque rupture. Various animal models of oxidative stress support the notion that ROS have a causal role in atherosclerosis and other cardiovascular diseases. Human investigations also support the oxidative stress hypothesis of atherosclerosis. Oxidative stress is the unifying mechanism for many CVD risk factors, which additionally supports its central role in CVD. Despite the demonstrated role of antioxidants in cellular and animal studies, the ineffectiveness of antioxidants in reducing cardiovascular death and morbidity in clinical trials has led many investigators to question the importance of oxidative stress in human atherosclerosis. Others have argued that the prime factor for the mixed outcomes from using antioxidants to prevent CVD may be the lack of specific and sensitive biomarkers by which to assess the oxidative stress phenotypes underlying CVD. A better understanding of the complexity of cellular redox reactions, development of a new class of antioxidants targeted to specific subcellular locales, and the phenotype-genotype linkage analysis for oxidative stress will likely be avenues for future research in this area as we move toward the broader use of pharmacological and regenerative therapies in the treatment and prevention of CVD.

    Antioxidants are ineffective in reducing cardiovascular death despite evidence for oxidative stress (OS) in cardiovascular diseases (CVD) from animal and human investigations. A better understanding of cellular redox reactions, development of targeted antioxidants and phenotype-genotype linkage analysis for OS are necessary for broader use of pharmacological and regenerative therapies for CVD.

    Key Words: reactive oxygen species ? NAD(P)H oxidase ? mitochondria ? atherosclerosis ? antioxidants

    Introduction

    Cardiovascular diseases (CVD)—coronary artery disease, hypertension, congestive heart failure, and stroke—are the leading cause of death and disability in the Western world.1 In the United States, the CVD death toll is nearly one million each year, and in 2002 the estimated cost of CVD treatment was $326.6 billion.2 To provide early prognosis and better therapies for preventing and curing these diseases, an understanding of the basic pathophysiologic mechanisms of CVD is essential.

    Atherosclerosis: An Overview

    The majority of cardiovascular disease results from complications of atherosclerosis. An important initiating event for atherosclerosis may well be the transport of oxidized low-density lipoprotein (Ox-LDL) across the endothelium into the artery wall.3 This is likely to occur at the sites of endothelial damage which are caused by Ox-LDL itself as well as physical or chemical forces and infection (Figure 1).4 Endothelial cells, smooth muscle cells (SMCs), and macrophages are the sources of oxidants for the oxidative modification of phospholipids. Ox-LDL can damage endothelial cells and induce the expression of adhesion molecules such as P-selectin5 and chemotactic factors such as monocyte chemoattractant protein-1 and macrophage colony stimulating factor (CSF).6,7 These processes lead to the tethering, activation, and attachment of monocytes and T lymphocytes to the endothelial cells.8 Endothelial cells, leukocytes, and SMCs then secrete growth factors and chemoattractants which effect the migration of monocytes and leukocytes into the subendothelial space.9 Monocytes ingest lipoproteins and morph into macrophages; macrophages generate reactive oxygen species (ROS), which convert Ox-LDL into highly oxidized LDL, which is, in turn, taken up by macrophages to form foam cells. Foam cells combine with leukocytes to become the fatty streak, and as the process continues foam cells secrete growth factors that induce SMC migration into the intima. SMC proliferation, coupled with the continuous influx and propagation of monocytes and macrophages, converts fatty streaks to more advanced lesions and ultimately to a fibrous plaque that will protrude into the arterial lumen. Later, calcification can occur and fibrosis continues, yielding a fibrous cap that surrounds a lipid-rich core. This formation may also contain dead or dying SMCs. In acute coronary syndromes (eg, myocardial infarction), when fibrous plaques rupture (Figure 1), the formation and release of thrombi may ultimately occlude vessels.

    Figure 1. Development of atherosclerosis. ROS produced by endothelial cells, SMCs, and macrophages oxidize LDL in the subendothelial space, at the sites of endothelial damage, initiating events that culminate in the formation of a fibrous plaque. Rupture of fibrous plaque leads to thrombus formation and occlusion of the vessel.

    Oxidative Stress and Atherosclerosis

    Oxygen is an abundant molecule in biological systems. Despite being a radical, it is sparingly reactive because its two unpaired electrons are situated in different molecular orbitals and demonstrate parallel spins. Thus, oxygen undergoes univalent reduction to form superoxide (O2) by means of enzymes such as the nicotinamide adenine dinucleotide (phosphate) (NADH/NAD(P)H) oxidases and xanthine oxidases (XO) (Figure 2). Nonenzymatically, oxygen can also become O2 by reacting with redox active compounds such as semiubiquinone of the mitochondrial electron transport chain.10 Superoxide anion is dismutated enzymatically to become hydrogen peroxide (H2O2) through the action of superoxide dismutases (SODs).11 In biological tissues, O2 can also undergo nonenzymatic transformation into H2O2 and singlet oxygen (1O2).12 H2O2 can react with other radicals such as transition metal Fe2+ to produce highly reactive hydroxyl radicals (OH·); this is known as the Fenton reaction (Figure 2). These radicals are capable of destroying biomolecules through oxidation. When Fe3+ initially oxidizes O2, molecular oxygen and Fe2+ are generated; the Fe2+ initiates the Fenton reaction and this regenerates Fe3+ which perpetuates the production of OH.13 Myeloperoxidase, a heme protein secreted by phagocytes, can amplify the oxidative potential of H2O2.14 At physiological concentrations of Cl–, hypochlorous acid (HOCl) is the major oxidant generated by the myeloperoxidase–H2O2--Cl– system,15 and HOCl can react with O2 to produce OH·.16 (Figure 2)

    Figure 2. Sources of ROS in vascular cells. Activated NAD(P)H oxidase, 12/15-LO and XO generate superoxide (O2). NOS switches from a coupled state to an uncoupled state and generates O2 with decreased availability of 5,6,7,8-tetrahydrobiopterin (BH4) or L-arginine. Dysfunctional mitochondrial respiratory chain is another source of O2 generation. SOD isoforms, Mn SOD, and CuZn SOD dismutate O2 to produce hydrogen peroxide (H2O2). Myeloperoxidase generates HOCl from H2O2 in the presence of Cl–. H2O2 reacts with transition metals to produce hydroxyl radicals (OH·).

    In response to growth factors and cytokines, and during normal metabolic events such as respiration and phagocytosis, eukaryotic cells produce oxidants. To compensate for this, the cells have evolved both enzymatic and nonenzymatic mechanisms to protect against oxidants’ toxic effects. The enzymatic mechanisms include the actions of enzymes such as SOD, catalase, and glutathione peroxidase. The nonenzymatic antioxidants include glutathione, ascorbate, and -tocopherol. However, in pathophysiologic circumstances, an excess of oxidants can overwhelm the scavenging capacity of cellular antioxidant systems. The subsequent oxidative stress damages the cell’s lipids, membranes, proteins, and DNA.

    NAD(P)H Oxidase

    Recent evidence indicates that membrane-bound NAD(P)H oxidases are the major source of O2 generation, and both NAD(P)H oxidase–derived O2 and mitochondrial-derived O2 constitute the bulk of this radical in the vasculature.17,18 However, the vascular-derived NAD(P)H oxidase has a similar but distinct structure from the phagocytic enzyme. The phagocytic oxidase contains the membrane-bound subunits gp91phox (Nox2) and p22phox, the catalytic site of the oxidase and the cytosolic components p47phox, p67phox, and G-protein rac1 or rac2.19 Endothelial cells and adventitial fibroblasts possess all the components of phagocytic oxidase.20,21 Vascular smooth muscle cells (VSMCs) have gp91phox homologues: NAD(P)H oxidase subunits Nox1 and Nox4.22,23,24 The most important distinction between phagocytic cells and VSMCs is that the latter do not appear to possess p67phox.25 The comprehensive term "vascular NAD(P)H oxidase" will be used henceforth to include the NAD(P)H oxidases present in VSMCs, endothelial cells, and adventitial fibroblasts. Vascular oxidase is active during normal metabolism, and a sustained activation of this enzyme occurs in response to agonists. Also, vascular oxidase produces much less superoxide (by several orders of magnitude) than does phagocytic oxidase.26 One of the early steps in the activation of the NAD(P)H oxidase is the phosphorylation and translocation of p47phox.27

    Cell culture studies and animal models provide evidence for the critical role of NAD(P)H oxidase–derived oxidative stress in atherosclerosis. NAD(P)H oxidase and O2 production are increased in vascular cells by a number of agonists associated with the pathogenesis. These agonists include angiotensin II (Ang II), thrombin, platelet-derived growth factor (PDGF), and tumor necrosis factor (TNF)-.25,28–30 Depending on vascular flow conditions, temporal regulation of NAD(P)H oxidase also occurs. Cell culture studies have shown that the atheroprotective laminar shear stress downregulates NAD(P)H oxidase activity, whereas atherogenic oscillatory shear stress induces a sustained increase in the oxidase activity.31 In hypercholesterolemic rabbits prone to atherosclerosis, enhanced AT1 receptor regulation and increased NADH-dependent vascular O2 production were associated with endothelial dysfunction.32 AT1 receptor antagonists not only inhibited the oxidase and improved endothelial function but also reduced early plaque formation, suggesting that oxidative stress plays a central role in the early stages of atherosclerosis. Recently, using NAD(P)H oxidase–deficient mice, we reported evidence for this.33 Mice lacking the p47phox gene not only had lower levels of aortic O2 production compared with wild-type mice, but also, when in a hypercholesterolemic apolipoprotein E–deficient (apoE(–/–)) background, had significantly fewer lesions in their descending aortas compared with apoE (–/–) mice. Further support for NAD(P)H oxidase–induced oxidative stress in neointimal hyperplasia comes from studies of balloon-injured porcine34 and rat35 coronary arteries. In the rat balloon injury model, Nox1 and p22phox were upregulated for 3 to 15 days, gp91phox for 7 to 15 days, and Nox4 at only 15 days after injury, suggesting not only a spatiotemporal but also a dynamic and differential regulation of the NAD(P)H-oxidase isoforms. In the rat carotid artery model, use of the gp91 ds-Tat peptide, a chimeric peptide consisting of a fragment from the Tat peptide of the HIV virus plus a fragment of gp91phox that prevents the interaction of p47phox with Nox subunits in cell free systems, provided direct evidence for the role of this oxidase in angioplasty-induced neointimal hyperplasia.36 The Tat fragment allowed the ready uptake of the peptide inhibitor into the cells and blocked both ROS production and neointima/media area and thickness. After angioplasty in these animals, the Tat peptide also inhibited peroxynitrite formation and stretch-induced ROS generation in distended vessels. Together these data clearly show that NAD(P)H oxidase–induced oxidative stress is causal in atherosclerosis.

    In human studies there is also considerable evidence to show the importance of NAD(P)H oxidase–derived oxidative stress in atherosclerosis. High levels of vasoactive agonists that induce oxidative stress in vitro were observed in human atherosclerotic plaques. Ang II was observed at the shoulder regions—the presumed rupture site of atherosclerotic plaques in human coronary arteries—in acute myocardial infarction; Ang II also colocalizes with the AT1 receptor.37 Similarly, increased expression of the thrombin receptor was observed in human atheroma.38 Recently, Azumi et al39 reported that in atherosclerotic human coronary arteries ROS production and Ox-LDL are spatially associated with NAD(P)H oxidase subunit p22phox. This suggests that ROS catalyze the formation of Ox-LDL, which leads to its uptake by macrophages and thus results in forming activated foam cells.40 These investigators also reported that ROS production was significantly higher in unstable versus stable angina pectoris, which suggests that ROS might also modulate plaque stability. These results are further supported by the observation that intimal SMCs, but not medial SMCs and macrophages, express high levels of NAD(P)H oxidase subunits.41 Plaque instability is initiated when ROS induce the expression of matrix-degrading enzymes such as matrix metalloproteinase (MMP)-2 and MMP-9.42 ROS also induce apoptosis in endothelial and SMCs through Ox-LDL.39

    Other Oxidase Systems

    XO generates O2 by catalyzing hypoxanthine and xanthine to uric acid. Under pathophysiologic conditions, this is another major source of vascular oxidative stress.10 Xanthine oxidase exists in plasma and endothelial cells but not in SMCs.18 In hypercholesterolemic rabbits, atherosclerosis resulting from diet was ascribed to XO activation–induced oxidative stress.43 In hypercholesterolemic patients, vasodilation is improved by using the XO inhibitor oxypurinol. The role for XO in atherosclerosis is further corroborated by the following observations:44

    In the coronary arteries of patients with coronary artery disease (CAD), electron spin resonance studies show significant activation of both NAD(P)H oxidase and XO;

    In these same patients, endothelial XO is inversely proportional and positively related to the effect of vitamin C on endothelium-dependent vasodilation; and

    In asymptomatic young individuals with familial hypercholesterolemia, the increase of vascular XO activity is an early event.

    In addition, lipoxygenases are another important source of ROS production in the vascular wall; these non-heme containing dioxygenases oxidize polyunsaturated fatty acids to hydroperoxy fatty-acid derivatives.10 Leukocyte-type 12/15-lipoxygenase (LO) and its products, 12(S)-hydroxyeicosatetraenoic acid[12(s)-HETE)] and 15(S)-HETE, are also implicated in atherogenesis. In addition, homozygous deletion of the 12/15-LO gene caused remarkable inhibition of early atherosclerosis in apoE-deficient mice.45 Inhibiting 12/15-LO reduced blood pressure in hypertensive rats46 and blocked intimal hyperplasia in balloon-injured rat carotid arteries.47 These findings bolster support for the role of this enzyme in vascular pathology. Furthermore, 12/15-LO activation leads to SMC growth, hypertrophy, and inflammatory gene expression; SMCs deficient in the enzyme show reduced generation of mitogen-induced ROS.48,49

    Uncoupling of NOS

    Nitric oxide synthases (NOSs), and in particular endothelial NOS (eNOS), can be potential sources of O2 under certain pathophysiologic conditions.18 Under normal conditions, these enzymes transfer electrons from a heme group in the oxygenase domain to the substrate L-arginine to form L-citrulline and nitric oxide (NO); 5,6,7,8-tetrahydrobiopterin (BH4) serves as a cofactor in this process.50 If the availability of either BH4 or L-arginine decreases, eNOS switches from a coupled state (generates NO) to an uncoupled state (generates O2) because the electrons from the heme reduce oxygen to form O2 (Figure 2). Increased vascular O2 production not only alters endothelium-dependent vascular relaxation through interaction with NO, but the resultant peroxynitrite can also oxidize BH4. This causes a deficiency of BH4 and the pathogenic uncoupling of NOS.51 It has been further reported that NAD(P)H oxidases are necessary for BH4 oxidation; this process does not occur in p47phox (–/–) mice that are hypertensive.52 The majority of studies report that NO induction decreases atherosclerosis in hypercholesterolemic animals and improves vascular function in hypercholesterolemic or hyperhomocysteinemic humans (see Cooke and Sydow, 2003).53

    Mitochondrial ROS

    Mitochondria provide energy (adenosine triphosphate; ATP) to the cell through oxidative phosphorylation. Oxidative phosphorylation is the process by which ATP is formed as electrons are transferred from NADH or FADH2 (generated through the Krebs cycle) to molecular oxygen. This occurs through a series of electron transport carriers localized in the inner mitochondrial membrane. The electron transport carriers include: complex I (NADH-ubiquinone oxidoreductase), complex II (succinate-ubiquinone oxidoreductase), complex III (ubiquinol-cytochrome c reductase), and complex IV (cytochrome c oxidase) (Figure 3). The transfer of more than 98% of electrons by the electron transport carriers/chain is coupled with the production of ATP. Only 1% to 2% of electrons leak out to form O2, and this is scavenged by manganese SOD (MnSOD/SOD2). However, during mitochondrial oxidative phosphorylation under pathophysiologic conditions, the electron transport chain may become uncoupled, leading to increased O2 production.54

    Figure 3. Inhibition of O2 production in mitochondria by inhibiting electron transport chain complex II (1), uncoupling oxidative phosphorylation (2), overexpressing uncoupling protein-1, or using SOD (3) leads to blockage of hyperglycemia-induced damage in endothelial cells.

    Mitochondrial DNA (MtDNA) is prone to oxidative damage because of several factors. These include the proximity of MtDNA to the sources of ROS generation in the mitochondrial inner membrane, the lack of protective histone-like proteins, and poor DNA damage-repair activity.55 MtDNA damage eventually results in reduced mtRNA transcription and, therefore, a loss of function.56 In accordance with this, we have shown that exogenous ROS-mediated mtDNA damage decreases mtDNA-encoded gene transcription in a dose-dependent manner.57 We have also demonstrated that the extent of atherosclerosis correlated well with mtDNA damage in the aortas from humans and apoE (–/–) mice and preceded atherogenesis in young apoE (–/–) mice.58 Furthermore, apoE (–/–) mice deficient in SOD2 showed increased mtDNA damage at earlier stages as well as an accelerated atherogenesis phenotype at arterial branch points. Together, the above reports indicate that mitochondrial ROS are clearly associated with enhanced susceptibility to atherosclerosis. Future and clinical trials are imperative to ascertain whether mtDNA damage can be used as a diagnostic oxidative phenotypic marker for atherosclerosis.

    Oxidative Stress and Hypertension

    More than 50 million Americans experience systolic hypertension,59 and hypertension is a risk factor for many other vascular diseases including atherosclerosis and stroke. The molecular basis of hypertension is complex; more than 50 genes have been implicated in the regulation of blood pressure.60 However, in the recent past, the role of the AT1 receptor in regulating hypertension has been the subject of intense investigation in both in vitro and animal models. Ang II modulates hypertension through its effect on the renin-angiotensin system, and the stimulation of AT1 receptors in the vascular wall leads to activation of NADH/NAD(P)H oxidase in vascular cells. The resultant oxidative stress is considered a unifying mechanism for hypertension and atherosclerosis.61,62 In addition to its indirect effect on NAD(P)H oxidase activation through AT1 receptor stimulation, Ang II can also directly regulate NAD(P)H activation by inducing a rapid translocation of small GTPase rac1 to the cell membrane,63 or by phosphorylating and translocating p47phox to cell membrane (Figure 4).64 Mechanical stretch, a hallmark of arterial hypertension, was recently shown to induce p47phox membrane translocation and NAD(P)H oxidase activation in VSMC.24 Stretch-stimulated NAD(P)H oxidase activation was absent in p47phox (–/–) cells and might play an important role in hypertension-induced vessel wall remodeling through activation of MMP-2.

    Figure 4. Mechanical stretch and AT1 receptor activation induce p47phox and rac1 activation, membrane translocation, and NAD(P)H oxidase activation. ROS generation by activated NAD(P)H oxidase and 12/15-lipoxygenase leads to hypertension.

    The critical role for oxidative stress in hypertension was demonstrated in animal models: when spontaneously hypertensive rats were treated with statins, O2 production and AT1-receptor activation decreased with a concomitant reduction in blood pressure.65 As mentioned above, the chimeric gp91ds-Tat peptide that inhibits NAD(P)H oxidase activation also attenuated Ang II-induced ROS production and blood pressure increases in mice.66 Similarly, a hypertensive response to Ang II infusion was blunted in p47phox (–/–) mice.67 Transgenic mice that expressed constitutively active rac1 in VSMCs exhibited a hypertensive phenotype; increased ROS production and treatment with antioxidants reversed hypertension.68 Oxidative stress induced by elevated 12-LO has also been implicated in hypertension in rat models. This suggests that multiple oxidative mechanisms are involved in hypertension.69 Consistent with this, increased levels of 12-LO and its product, 12-hydroperoxyeicosatetraenoic acid [12-(S)-HETE], were observed in patients with essential hypertension.70 Treatment of hypertensive patients with AT1 receptor blockers not only reduced blood pressure and diminished levels of malondialdehyde (MDA), a marker of oxidative stress,71 but also reversed hypertension-induced wall-structural changes in resistance arteries.72 Together, these observations strongly suggest that oxidative stress is a modulator of hypertension, a risk factor for atherosclerosis.

    Oxidative Stress and Heart Failure

    The importance of oxidative stress in chronic heart failure can be gauged by the fact that antioxidants prevent the progression of several pathological processes—such as cardiac hypertrophy, cardiac myocyte apoptosis, ischemia-reperfusion, and myocardial stunning—which lead to heart failure in animal models.73 In rat cardiac myocytes, TNF- and Ang II induced hypertrophy in a ROS-dependent manner; antioxidant use, including butylated hydroxyanisole, vitamin E, and catalase, prevented this.74 Overexpression of catalase significantly reduced Ang II–induced hypertrophy, and transfection with antisense p22phox inhibited Ang II–induced H2O2 production. This suggests that NAD(P)H oxidase–induced oxidative stress led to the hypertrophy.75 During compensated hypertrophy in a guinea pig model, NAD(P)H oxidase–dependent ROS production significantly and progressively increased to a peak at the level of decompensated heart failure. This indicates that ROS may be important mediators of heart failure.76 Other sequelae in the failing heart which may result from XO77 and mitochondria78 include structural damage and contractile dysfunction. Increased production of ROS may decrease NO bioavailability and impair diastolic function.79 In addition, increased peroxynitrite may cause cytokine-induced myocardial contractile failure by inactivating sarcoplasmic Ca2+-ATPase and dysregulating Ca2+ homeostasis.80,81

    Emerging evidence demonstrates that oxidative stress in general and NAD(P)H oxidase–derived ROS in particular are important in human cardiac failure. In the failing myocardium of patients with ischemic or dilated cardiomyopathy, NAD(P)H oxidase–derived ROS were upregulated.82 Plasma TNF- levels and platelet-derived NAD(P)H oxidase activity were also elevated in patients with heart failure.83 In addition, NAD(P)H oxidase activation and increased translocation of regulatory p47phox from the cytosol to the sarcolemmal membrane were recently observed in failing human myocardium.84 These combined results suggest that oxidative stress has a role in the pathophysiologic cardiac dysfunction in heart failure.

    Oxidative Stress and Stroke

    Results from a number of studies implicate oxidative stress in brain injury after ischemia and reperfusion. ROS produced during cerebral ischemia induce lipid peroxidation, protein oxidation, and DNA damage.85,86 In rodents, manipulating the genes responsible for antioxidant enzyme production provided the clearest evidence for oxidative stress in ischemia-induced brain injury. In SOD2 knockout mice, exacerbated infarct size and upregulated oxidative stress markers such as increased mitochondrial cytochrome c release and DNA fragmentation were observed after permanent focal cerebral ischemia.87,88 In contrast, mice that overexpress SOD2 showed neuronal protection.89 Similarly, a reduction in blood-brain barrier disruption and infarct size along with decreased oxidative DNA damage and DNA fragmentation were observed after photothrombotic ischemia in copper/zinc SOD (SOD1) transgenic mice as compared with wild-type mice.90 In a mouse ischemia-reperfusion model, deficiency of the antioxidant enzyme GPx-1 resulted in a threefold increase in brain infarct volume, early activation of caspase-3 expression, and enhanced apoptosis when compared with the wild-type mouse.91 Consistent with this, infusion of ebselen, a GPx mimic, before and during middle cerebral artery occlusion in rats conferred significant protection against ischemic damage.92 Support for the role of oxidative stress in cerebral ischemia was also obtained from mice deficient in NOS isoforms. In separate studies of mice deficient for neuronal (nNOS) and inducible NOS (iNOS) isoforms, reduced infarct size was observed after permanent focal cerebral ischemia.93–95 In endothelial-type NOS isoform (eNOS) knockout mice, increases in lesion volume were observed.96 In the brains of stroke-prone spontaneously hypertensive rats, electron spin resonance spectroscopy revealed high oxidative stress when compared with Wistar-Kyoto rats.97 Glutathione concentrations decreased before cerebral infarction in severe transient focal cerebral ischemia in rats which suggests that early oxidative stress might contribute to cerebral damage in stroke.98

    Significant increases in plasma homocystine, lipid peroxide, and NO, and decreased ascorbate levels were observed in stroke patients as compared with healthy controls.99 Also compared with healthy individuals, the circulating phagocytes of patients with ischemic stroke showed increased ROS production through mechanisms that were both opsonin receptor–dependent and –independent.100 These observations offer cumulative evidence that oxidative stress plays a distinct role in the pathogenesis of ischemic brain injury.

    CVD and Oxidative Stress Risk Factors

    Diabetes is a risk factor for atherosclerosis and, like atherosclerosis, it is progressive and associated with enhanced oxidative stress. In diabetic patients, enhanced oxidative stress in hyperglycemia is indicated by increased levels of the urinary isoprostane 8-epi-PGF2; with vitamin E treatment, it decreased significantly.101 In the recently concluded Diabetes Control and Complications Trial, a 6-year follow-up study revealed a significant increase in intima-media thickness (IMT) in diabetic patients compared with healthy subjects. After six years, intensive diabetes treatment decreased IMT progression, which clearly underlines the association of diabetes and atherosclerosis.102

    Both insulin-dependent diabetes mellitus (type 1) and noninsulin-dependent (type 2) diabetes are characterized by hyperglycemia.10 Hyperglycemia can induce oxidative stress by several different mechanisms. Autoxidation of glucose and the nonenzymatic glycation of proteins generate O2.103 Also, hyperglycemia induced by intra-arterial injection of dextrose impaired endothelium-dependent vasodilation, which was restored in nondiabetic subjects when they were treated with vitamin C.104 Glucose can also directly react with LDL phospholipids and apolipoprotein B (apoB) lysine groups to form the advanced glycation end products (AGEs) that facilitate lipid peroxidation.105 Consistent with this, an increased concentration of AGEs was correlated with endothelial dysfunction in patients with type 2 diabetes.106 In cultured bovine aortic endothelial cells, inhibiting elevated mitochondrial O2 blocked hyperglycemia-induced damage.107 Inhibition was accomplished through several pathway-specific means: (1) inhibiting electron transport chain complex II; (2) uncoupling oxidative phosphorylation; and (3) overexpression of uncoupling protein-1 or SOD2 (Figure 3). This suggests that in diabetes, mitochondrial ROS act through several independent pathways. Depletion of BH4, an essential cofactor for NOS, or L-arginine, a precursor of NOS, impaired endothelium-dependent vasorelaxation in the aortic rings of streptozotocin-induced diabetic rats.108 Similarly, endothelium-dependent vasodilation that had been attenuated in response to acetylcholine was improved in type 2 diabetic patients treated with BH4.109

    As in atherosclerosis, activation of NAD(P)H oxidase and enhanced O2 production were observed in the aortas of streptozotocin-induced diabetic rats.110 Inhibition of NOS with NG-nitro-L-arginine increased O2 levels in control vessels but reduced levels in diabetic vessels, which identifies NOS as a O2 source. As described earlier, NOS (NOSI and NOSIII) may become uncoupled in the presence of low levels of L-arginine or BH4, thus leading to the generation of O2 instead of NO. Depletion of L-arginine occurs in dyslipidemic conditions because Ox-LDL can inhibit endothelial L-arginine uptake.111 In contrast, peroxynitrite produced from the NO/O2 reaction can oxidize BH4 to inactive dihydrobiopterin.112 Hyperglycemia might also induce the dedifferentiation of aortic SMCs, which then produce more O2 and quench NO even when NO synthetic pathways are enhanced.113 Atherosclerotic lesions are initiated at reduced or disturbed oscillatory shear stress points in the arterial tree; reduced coronary shear stress was observed in diabetic patients compared with age-matched controls,114 clearly indicating that uncontrolled hyperglycemia leads to atherosclerosis.

    Other risk factors for vascular disease, such as obesity and cigarette smoking, are associated with systemic oxidative stress. In dyslipidemic obese mice, impaired high-density lipoprotein (HDL) defense and increased LDL oxidation were associated with increased atherosclerosis.115 The HDL-associated enzymes paraoxonase (PON1) and lecithin:cholesterol acyltransferase (LCAT) inhibit the oxidation of LDL.115,116 Transient overexpression of LCAT has led to a decrease in oxidative stress as measured by both a reduction in autoantibodies against MDA-modified LDL and a decrease in atherosclerosis. PON1 knockout mice were susceptible to atherosclerosis and exhibited enhanced expression of oxidative stress–related genes.117 Significantly lower levels of PON1 activity and concentration were observed in CAD patients compared with control subjects, which strongly suggests a role for PON1 in dyslipidemia-associated atherosclerosis.118

    A correlation between increasing body-mass index and enhanced systemic oxidative stress, as measured by urinary F2-isoprostanes, was observed in 3000 patients involved in the Framingham Heart Study.119 In addition, android obesity is also associated with elevated urinary F2-isoprostane levels.120 Cigarette smoke is replete with a number of oxidants,121 and consistent with this, smokers have increased F2-isoprostane levels in both plasma and urine.122 In animal models, brief exposure to cigarette smoke promoted atherosclerotic plaque development,123 and the combination of secondhand smoke and hypercholesterolemia resulted in a greater extent of mtDNA damage and atherosclerosis.124 In addition, benzo[a]pyrene, a polycyclic aromatic hydrocarbon present in tobacco smoke, can modulate VSMCs from a quiescent to a proliferative phenotype both in vitro and in vivo.125,126

    Antioxidants and Vascular Diseases

    Despite the preponderance of evidence for the association of increased oxidative stress with various vascular diseases, using antioxidants to prevent CVD has produced mixed outcomes.127,128 Of the 12 studies that used antioxidant vitamins at varying concentrations and follow-up times, 5 showed benefit with regard to their respective primary end points. In the Cambridge Heart Antioxidant Study (CHAOS), natural -tocopherol (RRR-AT) at a dose of either 400 or 800 IU/d caused a significant reduction in the combined primary end point of cardiovascular death and nonfatal myocardial infarction.129 In the Secondary Prevention with Antioxidants of Cardiovascular disease in End-stage renal disease (SPACE) study, administering 800 IU/d RRR-AT to hemodialysis patients with preexisting cardiovascular disease significantly reduced the composite primary end point, which included fatal and nonfatal myocardial infarction; ischemic stroke; peripheral vascular disease; and unstable angina.130 An investigation of transplant-associated atherosclerosis with a small sample size (total N=40) revealed that progression of the coronary intimal index (plaque area/vessel area) was inhibited with the combined supplementation of RRR-AT (800 IU/d) and ascorbic acid (AA: 1000 mg/d).131 In the Antioxidant Supplementation in Atherosclerosis Prevention (ASAP) study (N=440), a combination of RRR-AT (272 IU/d) and slow-release AA (500 mg/d) significantly decreased carotid intima-media thickness in hypercholesterolemic males.132 During 16 years of follow-up with the women in the Nurses’ Health Study (N=85,118), vitamin C intake of >359 mg/d from diet plus supplements or supplement use alone was associated with a significant reduction in nonfatal and fatal myocardial infarction.133 Contrary to these positive study outcomes, 7 other antioxidant supplementation studies did not show any effect on the primary end points of cardiovascular events (reviewed in Jialal and Devaraj, 2003).127

    The apparent inability for antioxidants to prevent CVD may be attributable to several reasons. One reason is that although antioxidant supplementation can be effective with a proved methodology for oxidative stress, we are still unable to deduce a priori those subjects who will be responders or nonresponders. Another reason for antioxidant ineffectiveness could relate to the optimum dose and type of antioxidants being used; efficacious threshold doses have been suggested at 800 IU/d of RRR-AT and 500 mg/d of AA.127,134 However, a key issue is antioxidant formulation; 5 of 7 antioxidant supplementation trials that reported inefficacy on primary end points used all-racemic-AT (all-rac-AT), but 4 trials described above with successful results used RRR-AT.127 The third reason why antioxidants are not yet successful in CVD prevention could be the complexity of redox reactions in vivo and the potential for a paradoxical increase in oxidant generation by antioxidants themselves. For instance, vitamin C supplementation exerted prooxidant as well as antioxidant effects in healthy volunteers; however, high doses of this antioxidant increased DNA damage.135 It was hypothesized that transition metal ions, released from metalloproteins after initial oxidative stress, act as catalysts in the presence of antioxidants to exacerbate the free radical-induced damage.136

    Ongoing research to further elucidate the connection between oxidative stress and CVD is necessary to distinguish the effective use of antioxidant vitamins for the primary prevention of vascular pathologies. In the meantime, we must continue to encourage patients to adopt healthy diets and lifestyles, and to carefully adhere to pharmacological regimens with such agents as statins and ACE inhibitors, to best protect themselves from the potential for vascular disease.

    Conclusions

    In conclusion, growing evidence from investigations of animal models and correlative data from human studies implicate oxidative stress in the development of CVD. However, a better understanding of the ROS-dependent signal transduction mechanisms, their localization, and the integration of both ROS-dependent transcriptional and signaling pathways in vascular pathophysiology is a prerequisite for effective pharmacological interventions for CVD. The development of a new class of antioxidants that are targeted to specific subcellular compartments such as mitochondria may help in combating CVD. Another important step toward effective treatment will be the development of sensitive and specific biomarkers that can be used clinically to assess the oxidative stress phenotypes that underlie various vascular pathologies. Ultimately, a phenotype–genotype linkage analysis that takes advantage of the recent advances in mouse and human genetics will be of immense help in the prognosis of people at risk for CVD. This should be complemented by studies that phenotype vascular cells and their progenitors, which will further contribute to the development of regenerative therapies. For example, regenerative therapies with mitochondria-rich endothelial progenitor cells may enhance myocardial activity and angiogenesis. Ultimately, remedial measures for oxidative stress–induced vascular malfunction will use a combination of preventive and regenerative therapies.

    Received September 20, 2004; accepted October 29, 2004.

    References

    Thom TJ. International mortality from heart disease: rates and trends. Int J Epidemiol. 1989; 18: S20–S28.

    Lefkowitz RJ, Willerson JT. Prospects for cardiovascular research. JAMA. 2001; 285: 581–587.

    Navab M, Berliner JA, Watson AD, Hama SY, Territo MC, Lusis AJ, Shih DM, Van Lenten BJ, Frank JS, Demer LL, Edwards PA, Fogelman AM. The Yin and Yang of oxidation in the development of the fatty streak. A review based on the 1994 George Lyman Duff Memorial Lecture. Arterioscler Thromb Vasc Biol. 1996; 16: 831–842.

    Ross R. Atherosclerosis: an inflammatory disease. N Engl J Med. 1999; 340: 115–126.

    Vora DK, Fang ZT, Liva SM, Tyner TR, Parhami F, Watson AD, Drake TA, Territo MC, Berliner JA. Induction of P-selectin by oxidized lipoproteins. Separate effects on synthesis and surface expression. Circ Res. 1997; 80: 810–818.

    Cushing SD, Berliner JA, Valente AJ, Territo MC, Navab M, Parhami F, Gerrity R, Schwartz CJ, Fogelman AM. Minimally modified low density lipoprotein induces monocyte chemotactic protein 1 in human endothelial cells and smooth muscle cells. Proc Natl Acad Sci U S A. 1990; 87: 5134–5138.

    Rajavashisth TB, Andalibi A, Territo MC, Berliner JA, Navab M, Fogelman AM, Lusis AJ. Induction of endothelial cell expression of granulocyte and macrophage colony-stimulating factors by modified low-density lipoproteins. Nature. 1990; 344: 254–257.

    McEver RP. Leukocyte-endothelial cell interactions. Curr Opin Cell Biol. 1992; 4: 840–849.

    Ross R. The pathogenesis of atherosclerosis: a perspective for the 1990s. Nature. 1993; 362: 801–809.

    Dr?ge W. Free radicals in the physiological control of cell function. Physiol Rev. 2002; 82: 47–95.

    Fridovich I. The biology of oxygen radicals. Science. 1978; 201: 875–880.

    Steinbeck MJ, Khan AU, Karnovsky MJ. Extracellular production of singlet oxygen by stimulated macrophages quantified using 9,10-diphenylanthracene and perylene in a polystyrene film. J Biol Chem. 1993; 268: 15649–15654.

    Kehrer JP. The Haber–Weiss reaction and mechanisms of toxicity. Toxicology. 2000; 149: 43–50.

    Schraufstatter IU, Browne K, Harris A, Hyslop PA, Jackson JH, Quehenberger O, Cochrane CG. Mechanisms of hypochlorite injury of target cells. J Clin Invest. 1990; 85: 554–562.

    Foote CS, Goyne TE, Lehrer RI. Assessment of chlorination by human neutrophils. Nature. 1983; 301: 715–716.

    Herdener M, Heigold S, Saran M, Bauer G. Target cell-derived superoxide anions cause efficiency and selectivity of intercellular induction of apoptosis. Free Radic Biol Med. 2000; 29: 1260–1271.

    Mohazzab KM, Kaminski PM, Wolin MS. NADH oxidoreductase is a major source of superoxide anion in bovine coronary artery endothelium. Am J Physiol. 1994; 266: H2568–H2572.

    Harrison D, Griendling KK, Landmesser U, Hornig B, Drexler H. Role of oxidative stress in atherosclerosis. Am J Cardiol. 2003; 91: 7A–11A.

    Babior BM. NADPH oxidase: an update. Blood. 1999; 93: 1464–1476.

    Jones SA, O’Donnell VB, Wood JD, Broughton JP, Hughes EJ, Jones OTG. Expression of phagocyte NAD(P)H oxidase components in human endothelial cells. Am J Physiol. 1996; 271: H1626–H1634.

    Pagano PJ, Clark JK, Cifuentes-Pagano ME, Clark SM, Callis GM, Quinn MT. Localization of a constitutively active, phagocyte-like NADPH oxidase in rabbit aortic adventitia: enhancement by angiotensin II. Proc Natl Acad Sci U S A. 1997; 94: 14483–14488.

    Suh YA, Arnold RS, Lassègue B, Shi J, Xu X, Sorescu D, Chung AB, Griendling KK, Lambeth JD. Cell transformation by the superoxide-generating oxidase Mox1. Nature. 1999; 401: 79–82.

    Lambeth J, Cheng G, Arnold R, Edens W. Novel homologs of gp91phox. Trends Biochem Sci. 2000; 25: 459–461.

    Grote K, Flach I, Luchtefeld M, Akin E, Holland SM, Drexler H, Schieffer B. Mechanical stretch enhances mRNA expression and proenzyme release of matrix metalloproteinase-2 (MMP-2) via NAD(P)H oxidase-derived reactive oxygen species. Circ Res. 2003; 92: e80–e86.

    Patterson C, Ruef J, Madamanchi NR, Barry-Lane P, Hu Z, Horaist C, Ballinger CA, Brasier AR, Bode C, Runge MS. Stimulation of vascular cell NAD(P)H oxidase by thrombin; evidence that p47phox may participate in forming this oxidase in vitro and in vivo. J Biol Chem. 1999; 274: 19814–19822.

    Griendling KK, Sorescu D, Ushio-Fukai M. NAD(P)H oxidase: role in cardiovascular biology and disease. Circ Res. 2000; 86: 494–501.

    Sellmayer A, Obermeier H, Danesch U, Aepfelbacher M, Weber PC. Arachidonic acid increases activation of NADPH oxidase in monocytic U937 cells by accelerated translocation of p47-phox and co-stimulation of protein kinase C. Cell Signal. 1996; 8: 397–402.

    Ushio-Fukai M, Zafari AM, Fukui T, Ishizaka N, Griendling KK. p22phox is a critical component of the superoxide-generating NADH/NADPH oxidase system and regulates angiotensin II-induced hypertrophy in vascular smooth muscle cells. J Biol Chem. 1996; 271: 23317–23321.

    Sundaresan M, Yu ZX, Ferrans VJ, Irani K, Finkel T. Requirement for generation of H2O2 for platelet-derived growth factor signal transduction. Science. 1995; 270: 296–299.

    De Keulenaer GW, Alexander RW, Ushio-Fukai M, Ishizaka N, Griendling KK. Tumor necrosis factor alpha activates a p22phox-based NADH oxidase in vascular smooth muscle cells. Biochem J. 1998; 329: 653–657.

    Hwang J, Ing MH, Salazar A, Lassegue B, Griendling K, Navab M, Sevanian A, Hsiai TK. Pulsatile versus oscillatory shear stress regulates NADPH oxidase subunit expression: implication for native LDL oxidation. Circ Res. 2003; 93: 1225–1232.

    Warnholtz A, Nickenig G, Schulz E, Macharzina R, Brasen JH, Skatchkov M, Heitzer T, Stasch JP, Griendling KK, Harrison DG, B?hm M, Meinertz T, Münzel T. Increased NADH-oxidase–mediated superoxide production in the early stages of atherosclerosis: evidence for involvement of the renin-angiotensin system. Circulation. 1999; 99: 2027–2033.

    Barry-Lane P, Patterson C, van der Merwe M, Hu Z, Holland S, Yeh E, Runge M. p47phox is required for atherosclerotic lesion progression in ApoE–/– mice. J Clin Invest. 2001; 108: 1513–1522.

    Shi Y, Niculescu R, Wang D, Patel S, Davenpeck KL, Zalewski A. Increased NAD(P)H oxidase and reactive oxygen species in coronary arteries after balloon injury. Arterioscler Thromb Vasc Biol. 2001; 21: 739–745.

    Sz?cs K, Lassègue B, Sorescu D, Hilenski LL, Valppu L, Couse TL, Wilcox JN, Quinn MT, Lambeth JD, Griendling KK. Upregulation of Nox-based NAD(P)H oxidases in restenosis after carotid injury. Arterioscler Thromb Vasc Biol. 2002; 22: 21–27.

    Jacobson GM, Dourron HM, Liu J, Carretero OA, Reddy DJ, Andrzejewski T, Pagano PJ. Novel NAD(P)H oxidase inhibitor suppresses angioplasty-induced superoxide and neointimal hyperplasia of rat carotid artery. Circ Res. 2003; 92: 637–643.

    Schieffer B, Schieffer E, Hilfiker-Kleiner D, Hilfiker A, Kovanen PT, Kaartinen M, Nussberger J, Harringer W, Drexler H. Expression of angiotensin II and interleukin 6 in human coronary atherosclerotic plaques: potential implications for inflammation and plaque instability. Circulation. 2000; 101: 1372–1378.

    Nelken NA, Soifer SJ, O’Keefe J, Vu TK, Charo IF, Coughlin SR. Thrombin receptor expression in normal and atherosclerotic human arteries. J Clin Invest. 1992; 90: 1614–1621.

    Azumi H, Inoue N, Ohashi Y, Terashima M, Mori T, Fujita H, Awano K, Kobayashi K, Maeda K, Hata K, Shinke T, Kobayashi S, Hirata K, Kawashima S, Itabe H, Hayashi Y, Imajoh-Ohmi S, Itoh H, Yokoyama M. Superoxide generation in directional coronary atherectomy specimens of patients with angina pectoris: important role of NAD(P)H oxidase. Arterioscler Thromb Vasc Biol. 2002; 22: 1838–1844.

    Channon KM. Oxidative stress and coronary plaque stability. Arterioscler Thromb Vasc Biol. 2002; 22: 1751–1752.

    Kalinina N, Agrotis A, Tararak E, Antropova Y, Kanellakis P, Ilyinskaya O, Quinn MT, Smirnov V, Bobik A. Cytochrome b558-dependent NAD(P)H oxidase-phox units in smooth muscle and macrophages of atherosclerotic lesions. Arterioscler Thromb Vasc Biol. 2002; 22: 2037–2043.

    Rajagopalan S, Meng XP, Ramasamy S, Harrison DG, Galis ZS. Reactive oxygen species produced by macrophage-derived foam cells regulate the activity of vascular matrix metalloproteinases in vitro. Implications for atherosclerotic plaque stability. J Clin Invest. 1996; 98: 2572–2579.

    Ohara Y, Peterson TE, Harrison DG. Hypercholesterolemia increases endothelial superoxide anion production. J Clin Invest. 1993; 91: 2546–2551.

    Spiekermann S, Landmesser U, Dikalov S, Bredt M, Gamez G, Tatge H, Reepschlager N, Hornig B, Drexler H, Harrison D. Electron spin resonance characterization of vascular xanthine and NAD(P)H oxidase activity in patients with coronary artery disease: relation to endothelium-dependent vasodilation. Circulation. 2003; 107: 1383–1389.

    Cyrus T, Witztum JL, Rader DJ, Tangirala R, Fazio S, Linton MF, Funk CD. Disruption of the 12/15-lipoxygenase gene diminishes atherosclerosis in apo E-deficient mice. J Clin Invest. 1999; 103: 1597–1604.

    Nozawa K, Tuck ML, Golub M, Eggena P, Nadler JL, Stern N. Inhibition of lipoxygenase pathway reduces blood pressure in renovascular hypertensive rats. Am J Physiol. 1990; 259: H1774–H1780.

    Gu JL, Pei H, Thomas L, Nadler JL, Rossi JJ, Lanting L, Natarajan R. Ribozyme-mediated inhibition of rat leukocyte-type 12-lipoxygenase prevents intimal hyperplasia in balloon-injured rat carotid arteries. Circulation. 2001; 103: 1446–1452.

    Reddy MA, Thimmalapura PR, Lanting L, Nadler JL, Fatima S, Natarajan R. The oxidized lipid and lipoxygenase product 12(S)-hydroxyeicosatetraenoic acid induces hypertrophy and fibronectin transcription in vascular smooth muscle cells via p38 MAPK and cAMP response element-binding protein activation. Mediation of angiotensin II effects. J Biol Chem. 2002; 277: 9920–9928.

    Reddy MA, Kim YS, Lanting L, Natarajan R. Reduced growth factor responses in vascular smooth muscle cells derived from 12/15-lipoxygenase-deficient mice. Hypertension. 2003; 41: 1294–1300.

    Vasquez-Vivar J, Kalyanaraman B, Martasek P, Hogg N, Masters BS, Karoui H, Tordo P, Pritchard KA Jr. Superoxide generation by endothelial nitric oxide synthase: the influence of cofactors. Proc Natl Acad Sci U S A. 1998; 95: 9220–9225.

    Laursen J, Somers M, Kurz S, McCann L, Warnholtz A, Freeman BA, Tarpey M, Fukai T, Harrison DG. Endothelial regulation of vasomotion in apoE-deficient mice: implications for interactions between peroxynitrite and tetrahydrobiopterin. Circulation. 2001; 103: 1282–1288.

    Landmesser U, Dikalov S, Price SR, McCann L, Fukai T, Holland SM, Mitch WE, Harrison DG. Oxidation of tetrahydrobiopterin leads to uncoupling of endothelial cell nitric oxide synthase in hypertension. J Clin Invest. 2003; 111: 1201–1209.

    Cooke J, Sydow K. A peculiar result and a fanciful hypothesis regarding L-arginine. Arterioscler Thromb Vasc Biol. 2003; 23: 1128.

    Boveris A, Cadenas E, Stoppani AO. Role of ubiquinone in the mitochondrial generation of hydrogen peroxide. Biochem J. 1976; 156: 435–444.

    Clayton D. Transcription of the mammalian mitochondrial genome. Annu Rev Biochem. 1984; 53: 573–594.

    Wallace D. Mitochondrial diseases in man and mouse. Science. 1999; 283: 1482–1488.

    Ballinger S, Patterson C, Yan CN, Doan R, Burow DL, Young CG, Yakes FM, Van Houten B, Ballinger CA, Freeman BA, Runge MS. Hydrogen peroxide- and peroxynitrite-induced mitochondrial DNA damage and dysfunction in vascular endothelial and smooth muscle cells. Circ Res. 2000; 86: 960–966.

    Ballinger S, Patterson C, Knight-Lozano CA, Burow DL, Conklin CA, Hu Z, Reuf J, Horaist C, Lebovitz R, Hunter GC, McIntyre K, Runge MS. Mitochondrial integrity and function in atherogenesis. Circulation. 2002; 106: 544–549.

    Chobanian AV, Bakris GL, Black HR, Cushman WC, Green LA, Izzo JLJ, Jones DW, Materson BJ, Oparil S, Wright JTJ, Roccella EJ, Joint National Committee on Prevention D, Evaluation, and Treatment of High Blood Pressure. National Heart, Lung, and Blood Institute;, Committee NHBPEPC. Seventh report of the Joint National Committee on Prevention, Detection, Evaluation, and Treatment of High Blood Pressure. Hypertension. 2003; 42: 1206–1252.

    Garbers DL, Dubois SK. The molecular basis of hypertension. Annu Rev Biochem. 1999; 68: 127–155.

    Zalba G, San Jose G, Moreno M, Fortuno M, Fortuno A, Beaumont F, Diez J. Oxidative stress in arterial hypertension: role of NAD(P)H oxidase. Hypertension. 2001; 38: 1395–1399.

    Nickenig G, Harrison DG. The AT(1)-type angiotensin receptor in oxidative stress and atherogenesis: part I: oxidative stress and atherogenesis. Circulation. 2002; 105: 393–396.

    Wassmann S, Laufs U, B?umer AT, Müller K, Konkol C, Sauer H, B?hm M, Nickenig G. Inhibition of geranylgeranylation reduces angiotensin II-mediated free radical production in vascular smooth muscle cells: involvement of angiotensin AT1 receptor expression and Rac1 GTPase. Mol Pharmacol. 2001; 59: 646–654.

    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: 981–987.

    Wassmann S, Laufs U, B?umer AT, Müller K, Ahlbory K, Linz W, Itter G, R?sen R, B?hm M, Nickenig G. HMG-CoA reductase inhibitors improve endothelial dysfunction in normocholesterolemic hypertension via reduced production of reactive oxygen species. Hypertension. 2001; 37: 1450–1457.

    Rey F, Cifuentes M, Kiarash A, Quinn M, Pagano P. Novel Competitive Inhibitor of NAD(P)H oxidase assembly attenuates vascular O(2)(-) and systolic blood pressure in mice. Circ Res. 2001; 89: 408–414.

    Landmesser U, Cai H, Dikalov S, McCann L, Hwang J, Jo H, Holland S, Harrison D. Role of p47(phox) in vascular oxidative stress and hypertension caused by angiotensin II. Hypertension. 2002; 40: 511–515.

    Gregg D, Rauscher FM, Goldschmidt-Clermont PJ. Rac regulates cardiovascular superoxide through diverse molecular interactions: more than a binary GTP switch. Am J Physiol Cell Physiol. 2003; 285: C723–734.

    DelliPizzi A, Guan H, Tong X, Takizawa H, Nasjletti A. Lipoxygenase-dependent mechanisms in hypertension. Clin Exp Hypertens. 2000; 22: 181–192.

    Gonzalez-Nunez D, Claria J, Rivera F, Poch E. Increased levels of 12(S)-HETE in patients with essential hypertension. Hypertension. 2001; 37: 334–338.

    Koh K, Ahn J, Han S, Kim D, Jin D, Kim H, Shin M, Ahn T, Choi I, Shin E. Pleiotropic effects of angiotensin II receptor blocker in hypertensive patients. J Am Coll Cardiol. 2003; 42: 905–910.

    Schiffrin EL, Park JB, Intengan HD, Touyz RM. Correction of arterial structure and endothelial dysfunction in human essential hypertension by the angiotensin receptor antagonist losartan. Circulation. 2000; 101: 1653–1659.

    Sorescu D, Griendling KK. Reactive oxygen species, mitochondria, and NAD(P)H oxidases in the development and progression of heart failure. Congest Heart Fail. 2002; 8: 132–140.

    Nakamura K, Fushimi K, Kouchi H, Mihara K, Miyazaki M, Ohe T, Namba M. Inhibitory effects of antioxidants on neonatal rat cardiac myocyte hypertrophy induced by tumor necrosis factor-alpha and angiotensin II. Circulation. 1998; 98: 794–799.

    Zafari A, Ushio-Fukai M, Akers M, Yin Q, Shah A, Harrison DG, Taylor WR, Griendling KK. Role of NADH/NADPH oxidase-derived H2O2 in angiotensin II-induced vascular hypertrophy. Hypertension. 1998; 32: 488–495.

    Li J, Gall NP, Grieve DJ, Chen M, Shah AM. Activation of NADPH oxidase during progression of cardiac hypertrophy to failure. Hypertension. 2002; 40: 477–484.

    Ekelund U, Harrison RW, Shokek O, Thakkar RN, Tunin RS, Senzaki H, Kass DA, Marban E, Hare JM. Intravenous allopurinol decreases myocardial oxygen consumption and increases mechanical efficiency in dogs with pacing-induced heart failure. Circ Res. 1999; 85: 437–445.

    Ide T, Tsutsui H, Kinugawa S, Utsumi H, Kang D, Hattori N, Uchida K, Arimura K, Egashira K, Takeshita A. Mitochondrial electron transport complex I is a potential source of oxygen free radicals in the failing myocardium. Circ Res. 1999; 85: 357–363.

    Grocott-Mason R, Anning P, Evans H, Lewis MJ, Shah AM. Modulation of left ventricular relaxation in isolated ejecting heart by endogenous nitric oxide. Am J Physiol. 1994; 267: H1804–H1813.

    Klebl B, Ayoub AT, Pette D. Protein oxidation, tyrosine nitration, and inactivation of sarcoplasmic reticulum Ca2+-ATPase in low-frequency stimulated rabbit muscle. FEBS Lett. 1998; 422: 381–384.

    Ferdinandy P, Danial H, Ambrus I, Rothery RA, Schulz R. Peroxynitrite is a major contributor to cytokine-induced myocardial contractile failure. Circ Res. 2000; 87: 241–247.

    Maack C, Kartes T, Kilter H, Schafers HJ, Nickenig G, Bohm M, Laufs U. Oxygen free radical release in human failing myocardium is associated with increased activity of rac1-GTPase and represents a target for statin treatment. Circulation. 2003; 108: 1567–1574.

    De Biase L, Pignatelli P, Lenti L, Tocci G, Piccioni F, Riondino S, Pulcinelli FM, Rubattu S, Volpe M, Violi F. Enhanced TNF alpha and oxidative stress in patients with heart failure: effect of TNF alpha on platelet O2- production. Thromb Haemost. 2003; 90: 317–325.

    Heymes C, Bendall JK, Ratajczak P, Cave AC, Samuel JL, Hasenfuss G, Shah AM. Increased myocardial NADPH oxidase activity in human heart failure. J Am Coll Cardiol. 2003; 41: 2164–2171.

    Traystman R, Kirsch JR, Koehler RC. Oxygen radical mechanisms of brain injury following ischemia and reperfusion. J Appl Physiol. 1991; 71: 1185–1195.

    Chan P. Reactive oxygen radicals in signaling and damage in the ischemic brain. J Cereb Blood Flow Metab. 2001; 21: 2–14.

    Murakami K, Kondo T, Kawase M, Li Y, Sato S, Chen SF, Chan PH. Mitochondrial susceptibility to oxidative stress exacerbates cerebral infarction that follows permanent focal cerebral ischemia in mutant mice with manganese superoxide dismutase deficiency. J Neurosci. 1998; 18: 205–213.

    Fujimura M, Morita-Fujimura Y, Kawase M, Copin JC, Calagui B, Epstein CJ, Chan PH. Manganese superoxide dismutase mediates the early release of mitochondrial cytochrome C and subsequent DNA fragmentation after permanent focal cerebral ischemia in mice. J Neurosci. 1999; 19: 3414–3422.

    Keller J, Kindy MS, Holtsberg FW, St Clair DK, Yen HC, Germeyer A, Steiner SM, Bruce-Keller AJ, Hutchins JB, Mattson MP. Mitochondrial manganese superoxide dismutase prevents neural apoptosis and reduces ischemic brain injury: suppression of peroxynitrite production, lipid peroxidation, and mitochondrial dysfunction. J Neurosci. 1998; 18: 687–697.

    Kim G, Lewen A, Copin J, Watson BD, Chan PH. The cytosolic antioxidant, copper/zinc superoxide dismutase, attenuates blood-brain barrier disruption and oxidative cellular injury after photothrombotic cortical ischemia in mice. Neuroscience. 2001; 105: 1007–1018.

    Crack P, Taylor JM, Flentjar NJ, de Haan J, Hertzog P, Iannello RC, Kola I. Increased infarct size and exacerbated apoptosis in the glutathione peroxidase-1 (Gpx-1) knockout mouse brain in response to ischemia/reperfusion injury. J Neurochem. 2001; 78: 1389–1399.

    Imai H, Graham DI, Masayasu H, Macrae IM. Antioxidant ebselen reduces oxidative damage in focal cerebral ischemia. Free Radic Biol Med. 2003; 34: 56–63.

    Huang Z, Huang PL, Panahian N, Dalkara T, Fishman MC, Moskowitz MA. Effects of cerebral ischemia in mice deficient in neuronal nitric oxide synthase. Science. 1994; 265: 1883–1885.

    Shimizu-Sasamata M, Bosque-Hamilton P, Huang PL, Moskowitz MA, Lo EH. Attenuated neurotransmitter release and spreading depression-like depolarizations after focal ischemia in mutant mice with disrupted type I nitric oxide synthase gene. J Neurosci. 1998; 18: 9564–9571.

    Iadecola C, Zhang F, Casey R, Nagayama M, Ross ME. Delayed reduction of ischemic brain injury and neurological deficits in mice lacking the inducible nitric oxide synthase gene. J Neurosci. 1997; 17: 9157–9164.

    Huang Z, Huang PL, Ma J, Meng W, Ayata C, Fishman MC, Moskowitz MA. Enlarged infarcts in endothelial nitric oxide synthase knockout mice are attenuated by nitro-L-arginine. J Cereb Blood Flow Metab. 1996; 16: 981–987.

    Miyazaki H, Shoji H, Lee MC. Measurement of oxidative stress in stroke-prone spontaneously hypertensive rat brain using in vivo electron spin resonance spectroscopy. Redox Rep. 2002; 7: 260–265.

    Lerouet D, Beray-Berthat V, Palmier B, Plotkine M, Margaill I. Changes in oxidative stress, iNOS activity and neutrophil infiltration in severe transient focal cerebral ischemia in rats. Brain Res. 2002; 958: 166–175.

    El Kossi M, Zakhary MM. Oxidative stress in the context of acute cerebrovascular stroke. Stroke. 2000; 31: 1889–1892.

    Alexandrova M, Bochev PG, Markova VI, Bechev BG, Popova MA, Danovska MP, Simeonova VK. Changes in phagocyte activity in patients with ischaemic stroke. Luminescence. 2001; 16: 357–365.

    Davi G, Ciabattoni G, Consoli A, Mezzetti A, Falco A, Santarone S, Pennese E, Vitacolonna E, Bucciarelli T, Costantini F, Capani F, Patrono C. In vivo formation of 8-iso-prostaglandin f2alpha and platelet activation in diabetes mellitus: effects of improved metabolic control and vitamin E supplementation. Circulation. 1999; 99: 224–229.

    Nathan DM, Lachin J, Cleary P, Orchard T, Brillon DJ, Backlund JY, O’Leary DH, Genuth S; Diabetes Control and Complications Trial; Epidemiology of Diabetes Interventions and Complications Research Group. Intensive diabetes therapy and carotid intima-media thickness in type 1 diabetes mellitus. N Engl J Med. 2003; 348: 2294–2303.

    Baynes JW. Role of oxidative stress in development of complications in diabetes. Diabetes. 1991; 40: 405–412.

    Beckman JA, Goldfine AB, Gordon MB, Creager MA. Ascorbate restores endothelium-dependent vasodilation impaired by acute hyperglycemia in humans. Circulation. 2001; 103: 1618–1623.

    Bucala R, Makita Z, Koschinsky T, Cerami A, Vlassara H. Lipid advanced glycosylation: pathway for lipid oxidation in vivo. Proc Natl Acad Sci U S A. 1993; 90: 6434–6438.

    Tan KC, Chow WS, Ai VH, Metz C, Bucala R, Lam KS. Advanced glycation end products and endothelial dysfunction in type 2 diabetes. Diabetes Care. 2002; 25: 1055–1059.

    Nishikawa T, Edelstein D, Du XL, Yamagishi S, Matsumura T, Kaneda Y, Yorek MA, Beebe D, Oates PJ, Hammes HP, Giardino I, Brownlee M. Normalizing mitochondrial superoxide production blocks three pathways of hyperglycaemic damage. Nature. 2000; 404: 787–790.

    Pieper GM. Acute amelioration of diabetic endothelial dysfunction with a derivative of the nitric oxide synthase cofactor, tetrahydrobiopterin. J Cardiovasc Pharmacol. 1997; 29: 8–15.

    Heitzer T, Krohn K, Albers S, Meinertz T. Tetrahydrobiopterin improves endothelium-dependent vasodilation by increasing nitric oxide activity in patients with Type II diabetes mellitus. Diabetologia. 2000; 43: 1435–1438.

    Hink U, Li H, Mollnau H, Oelze M, Matheis E, Hartmann M, Skatchkov M, Thaiss F, Stahl RAK, Warnholtz A, Meinertz T, Griendling K, Harrison DG, Forstermann U, Munzel T. Mechanisms underlying endothelial dysfunction in diabetes mellitus. Circ Res. 2001; 88: E14–E22.

    Vergnani L, Hatrik S, Ricci F, Passaro A, Manzoli N, Zuliani G, Brovkovych V, Fellin R, Malinski T. Effect of native and oxidized low-density lipoprotein on endothelial nitric oxide and superoxide production: key role of L-arginine availability. Circulation. 2000; 101: 1261–1266.

    Milstien S, Katusic Z. Oxidation of tetrahydrobiopterin by peroxynitrite: implications for vascular endothelial function. Biochem Biophys Res Commun. 1999; 263: 681–684.

    Pandolfi A, Grilli A, Cilli C, Patruno A, Giaccari A, Di Silvestre S, De Lutiis MA, Pellegrini G, Capani F, Consoli A, Felaco M. Phenotype modulation in cultures of vascular smooth muscle cells from diabetic rats: association with increased nitric oxide synthase expression and superoxide anion generation. J Cell Physiol. 2003; 196: 378–385.

    Irace C, Carallo C, Crescenzo A, Motti C, De Franceschi M, Mattioli P, Gnasso A. NIDDM is associated with lower wall shear stress of the common carotid artery. Diabetes. 1999; 48: 193–197.

    Mertens A, Verhamme P, Bielicki J, Phillips M, Quarck R, Verreth W, Stengel D, Ninio E, Navab M, Mackness B, Mackness M, Holvoet P. Increased low-density lipoprotein oxidation and impaired high-density lipoprotein antioxidant defense are associated with increased macrophage homing and atherosclerosis in dyslipidemic obese mice: LCAT gene transfer decreases atherosclerosis. Circulation. 2003; 107: 1640–1646.

    Rozenberg O, Rosenblat M, Coleman R, Shih DM, Aviram M. Paraoxonase (PON1) deficiency is associated with increased macrophage oxidative stress: studies in PON1-knockout mice. Free Radic Biol Med. 2003; 34: 774–784.

    Shih DM, Xia YR, Wang XP, Miller E, Castellani LW, Subbanagounder G, Cheroutre H, Faull KF, Berliner JA, Witztum JL, Lusis AJ. Combined serum paraoxonase knockout/apolipoprotein E knockout mice exhibit increased lipoprotein oxidation and atherosclerosis. J Biol Chem. 2000; 275: 17527–17535.

    Mackness B, Davies GK, Turkie W, Lee E, Roberts DH, Hill E, Roberts C, Durrington PN, Mackness MI. Paraoxonase status in coronary heart disease: are activity and concentration more important than genotype? Arterioscler Thromb Vasc Biol. 2001; 21: 1451–1457.

    Keaney JJ, Larson M, Vasan R, Wilson P, Lipinska I, Corey D, Massaro J, Sutherland P, Vita J, Benjamin E, Framingham S. Obesity and systemic oxidative stress: clinical correlates of oxidative stress in the Framingham Study. Arterioscler Thromb Vasc Biol. 2003; 23: 434–439.

    Davi G, Guagnano MT, Ciabattoni G, Basili S, Falco A, Marinopiccoli M, Nutini M, Sensi S, Patrono C. Platelet activation in obese women: role of inflammation and oxidant stress. JAMA. 2002; 288: 2008–2014.

    Frei B, Forte TM, Ames BN, Cross CE. Gas phase oxidants of cigarette smoke induce lipid peroxidation and changes in lipoprotein properties in human blood plasma. Protective effects of ascorbic acid. Biochem J. 1991; 277: 133–138.

    Morrow J, Frei B, Longmire A, Gaziano J, Lynch S, Shyr Y, Strauss W, Oates J, Roberts LI. Increase in circulating products of lipid peroxidation (F2-isoprostanes) in smokers. Smoking as a cause of oxidative damage. N Engl J Med. 1995; 332: 1198–1203.

    Penn A, Snyder CA. Inhalation of sidestream cigarette smoke accelerates development of arteriosclerotic plaques. Circulation. 1993; 88: 1820–1825.

    Knight-Lozano CA, Young CG, Burow DL, Hu ZY, Uyeminami D, Pinkerton KE, Ischiropoulos H, Ballinger SW. Cigarette smoke exposure and hypercholesterolemia increase mitochondrial damage in cardiovascular tissues. Circulation. 2002; 105: 849–854.

    Ramos KS, Zhang Y, Sadhu DN, Chapkin RS. The induction of proliferative vascular smooth muscle cell phenotypes by benzo(a)pyrene is characterized by up-regulation of inositol phospholipid metabolism and c-Ha-ras gene expression. Arch Biochem Biophys. 1996; 332: 213–222.

    Ou X, Ramos KS. Proliferative responses of quail aortic smooth muscle cells to benzopyrene: implications in PAH-induced atherogenesis. Toxicology. 1992; 74: 243–258.

    Jialal I, Devaraj S. Antioxidants and atherosclerosis: don’t throw out the baby with the bath water. Circulation. 2003; 107: 926–928.

    Patterson C, Madamanchi NR, Runge MS. The oxidative paradox: another piece in the puzzle. Circ Res. 2000; 87: 1074–1076.

    Stephens NG, Parsons A, Schofield PM, Kelly F, Cheeseman K, Mitchinson MJ. Randomised controlled trial of vitamin E in patients with coronary disease: Cambridge Heart Antioxidant Study (CHAOS). Lancet. 1996; 347: 781–786.

    Boaz M, Smetana S, Weinstein T, Matas Z, Gafter U, Iaina A, Knecht A, Weissgarten Y, Brunner D, Fainaru M, Green MS. Secondary prevention with antioxidants of cardiovascular disease in endstage renal disease (SPACE): randomised placebo-controlled trial. Lancet. 2000; 356: 1213–1218.

    Fang JC, Kinlay S, Beltrame J, Hikiti H, Wainstein M, Behrendt D, Suh J, Frei B, Mudge GH, Selwyn AP, Ganz P. Effect of vitamins C and E on progression of transplant-associated arteriosclerosis: a randomised trial. Lancet. 2002; 359: 1108–1113.

    Salonen RM, Nyyss?nen K, Kaikkonen J, Porkkala-Sarataho E, Voutilainen S, Rissanen TH, Tuomainen T-P, Valkonen V-P, Ristonmaa U, Lakka H-M, Vanharanta M, Salonen JT, Poulsen HE. Six-year effect of combined vitamin C and E supplementation on atherosclerotic progression: The Antioxidant Supplementation in Atherosclerosis Prevention (ASAP) study. Circulation. 2003; 107: 947–953.

    Osganian SK, Stampfer MJ, Rimm E, Spiegelman D, Hu FB, Manson JE, Willett WC. Vitamin C and risk of coronary heart disease in women. J Am Coll Cardiol. 2003; 42: 246–252.

    Frei B. To C or not to C, that is the question! J Am Coll Cardiol. 2003; 42: 253–255.

    Podmore ID, Griffiths HR, Herbert KE, Mistry N, Mistry P, Lunec J. Vitamin C exhibits pro-oxidant properties. Nature. 1998; 392: 559.

    Halliwell B. The antioxidant paradox. Lancet. 2000; 355: 1179–1180.(Nageswara R. Madamanchi; )