Abstract

Increased production of reactive oxygen species (ROS) has been implicated in the pathogenesis of cardiovascular diseases. Enzymatic systems such as the mitochondrial respiratory chain, vascular NAD(P)H oxidases, xanthine oxidase, and uncoupled endothelial nitric oxide synthase (eNOS) produce superoxide anion (O2.−) in vascular cells. While some O2.− rapidly degrades by reacting with nitric oxide (NO.), the O2.− signal preserved by dismutation into hydrogen peroxide (H2O2) exerts prolonged signaling effects. This review focuses on patterns and mechanisms whereby H2O2 modulates different aspects of endothelial cell function including endothelial cell growth and proliferation, endothelial apoptosis, endothelium-dependent vasorelaxation, endothelial cytoskeletal reorganization and barrier dysfunction, endothelial inflammatory responses, and endothelium-regulated vascular remodeling. These modulations of endothelial cell function may at least partially underlie H2O2 contribution to the development of vascular disease.

Increased production of reactive oxygen species (ROS) has been implicated in the pathogenesis of cardiovascular diseases such as atherosclerosis, restenosis, hypertension, diabetic vascular complications and heart failure [1–7]. One electron reduction of molecular oxygen forms superoxide anion (O2.−), which serves as a progenitor for hydrogen peroxide (H2O2). By rapidly inactivating nitric oxide (NO.), O2.− contributes to endothelial dysfunction [7]. H2O2 however modulates endothelial cell function via intricate mechanisms. Ambient production of O2.− and subsequently H2O2 at low levels, maintained by basal activity of pre-assembled endothelial NAD(P)H oxidases [3], or potential leakage from mitochondrial respiration [8], is necessary for endothelial cell growth and proliferation. Under pathological conditions, agonists-provoked activation of vascular NAD(P)H oxidases and subsequently activated xanthine oxidase or uncoupled eNOS produce H2O2 in large quantities [9–11], leading to detrimental consequences [12,13]. These dual-faced functions of H2O2 are similar to what is characteristic of NO.: vasodilating and protective at physiological concentrations but microbicidal and apoptotic when produced in large quantities. The current review discusses enzymes responsible for endothelial production of ROS, patterns and mechanisms whereby H2O2 influences different aspects of endothelial cell function and H2O2 contribution to vascular disease development.

1 Production and metabolism of hydrogen peroxide

In mammalian cells, potential enzymatic sources of ROS include the mitochondrial electron transport chain, the arachidonic acid metabolizing enzymes lipoxygenase and cycloxygenase, the cytochrome P450s, xanthine oxidase, NAD(P)H oxidases, uncoupled nitric oxide synthase (NOS), peroxidases, and other hemoproteins. These systems primarily catalyze one electron reduction of molecular oxygen to form O2.− which rapidly inactivates NO. to form peroxynitrite. Under ambient conditions, some O2.− is dismutated to H2O2 spontaneously or catalyzed by superoxide dismutase (SOD). Of interest, loss of NO. could lead to enhanced formation of H2O2. Some enzymes, such as xanthine oxidase and glucose oxidase, can directly produce H2O2 by donating two electrons to oxygen. In the presence of heavy metals, H2O2 undergoes Fenton reaction to form highly reactive hydroxyl radical (HO.). When bound to peroxidases such as catalase, H2O2 forms compound I which oxidizes NO. to nitrogen dioxide anion (NO2) and react with NO2 to form nitrogen dioxide radical (NO2.). NO2. in turn participates in nitrating events such as formation of nitrotyrosines [14,15]. Fig. 1 illustrates biochemical pathways of H2O2 generation and metabolism. When produced transiently at high concentrations, H2O2 and HO. are capable of directly oxidizing proteins and lipids, and causing DNA strand breaks. At relatively lower pathophysiological concentrations (nano- to micromolar range) however, H2O2 plays important signaling roles in endothelial cells, and these are the focus of the current review.

Biochemical pathways of hydrogen peroxide generation and metabolism. Molecular oxygen undergoes one or two-electron reduction (1 or 2e-) to form superoxide (O2.−) or hydrogen peroxide (H2O2) respectively. The majority of the bioactive H2O2 however, is derived from spontaneous or SOD (superoxide dismutase)-catalyzed (at the reaction speed of 2.0 × 109 mol/L−1.s−1) dismutation of O2.−. Degradation of H2O2 involves intracellular catalase (CAT), extracellular glutathione peroxidase (Gpx) or small molecules like thiols. Besides directly serving as a signaling intermediate, H2O2 also indirectly exerts its biological effects via metabolites such as hydroxyl radical (HO.) or compound I (product of H2O2 oxidation of Fe3+-containing enzymes such as myeloperoxidase, MPO). Of note, O2.− rapidly scavenges NO. at the reaction speed of 6.7 × 109 mol/L−1.s−1, representing one of the mechanisms whereby bioactive NO. diminishes independent of regulation of NO. synthase.
Fig. 1

Biochemical pathways of hydrogen peroxide generation and metabolism. Molecular oxygen undergoes one or two-electron reduction (1 or 2e-) to form superoxide (O2.−) or hydrogen peroxide (H2O2) respectively. The majority of the bioactive H2O2 however, is derived from spontaneous or SOD (superoxide dismutase)-catalyzed (at the reaction speed of 2.0 × 109 mol/L−1.s−1) dismutation of O2.−. Degradation of H2O2 involves intracellular catalase (CAT), extracellular glutathione peroxidase (Gpx) or small molecules like thiols. Besides directly serving as a signaling intermediate, H2O2 also indirectly exerts its biological effects via metabolites such as hydroxyl radical (HO.) or compound I (product of H2O2 oxidation of Fe3+-containing enzymes such as myeloperoxidase, MPO). Of note, O2.− rapidly scavenges NO. at the reaction speed of 6.7 × 109 mol/L−1.s−1, representing one of the mechanisms whereby bioactive NO. diminishes independent of regulation of NO. synthase.

2 Vascular NAD(P)H oxidases

Accumulating evidence has characterized a predominant role of vascular NAD(P)H oxidases in generating ROS in different vascular cells including endothelial cells, vascular smooth muscle and fibroblasts [1–6]. In endothelial cells, though there is much to be learned, it is clear that p47phox modulates enzymatic activity by interacting with membrane and cytosolic components to form the active complex. The gp91phox [also called Nox2, Nox for NAD(P)H oxidases, representing a family of novel NAD(P)H oxidases] is the catalytic core of this complex [16–19]. Studies using deficient mice or inhibitory peptide targeting Nox2 have established an essential role of Nox2 in producing ROS in endothelial cells [17,18,20,21]. Functions of the newly cloned gp91phox homologues; Nox1, Nox4 and Nox5 however, remain obscure. Nox4 seems to express more abundantly in endothelial cells compared to other Nox proteins, representing the major catalytic unit of the endothelial NAD(P)H oxidase responsive to growth halting [22,23]. Nox1 on the other hand, was upregulated by oscillatory shear stress, mediating ROS-dependent leukocyte adhesion to endothelium [24]. Furthermore, VEGF receptor-dependent activation of Nox1 induced angiogenic tube formation of endothelial cells [25]. The observations that Nox1 mediates growth signaling while Nox4 is growth suppressive in endothelial cells seems similar to what have been observed in vascular smooth muscle [26,27]. It seems odd that the same ROS-producing Nox proteins mediate different cellular responses. It is possible that the function of each Nox protein is dependent on its distinctive subcellular localization, and is subject to specific regulations by selective agonists [4]. For example, in vascular smooth muscle cells, Nox1 localizes to caveolae while Nox4 is found in focal adhesions [28]. In endothelial cells however, Nox4 was found at endoplasmic reticulum [23] whereas Nox2 is localized to peri-nuclear cytoskeletal structure [17].

Novel homologues of Nox-regulating proteins p47phox and gp67phox have been identified in epithelial cells (p41phox and p51phox respectively), serving as potent positive regulators for Nox1 [29–32]. In the same cells longer Nox proteins with peroxidase tails, namely Duox1 and Duox2, have been cloned [33,34] and shown to be functional [35,36]. These proteins are yet studied for their presence and function in endothelium and vascular smooth muscle. Of note, p22phox is the only other membrane component of the vascular NAD(P)H oxidases. Overexpression of p22phox led to stabilization of Nox1 and Nox4 proteins [37]. Recent studies have elegantly characterized physical and functional interactions between p22phox and Noxs (Nox1 and Nox4) in vascular smooth muscle [38,39]. Similar interactions might also occur in endothelial cells. In addition, a recent study reported that p22phox expression correlates well with expression of Nox4 in human arteries and that of Nox2 in veins [40].

In addition, vascular NAD(P)H oxidase-derived H2O2 is able to amplify ROS production in vascular cells. To date at least five different mechanisms are involved in this propagation, including enhanced intracellular iron uptake, and activation of sources of mitochondria, NAD(P)H oxidases, xanthine oxidase and uncoupled eNOS [1]. These feed-forward mechanisms form a vicious circle to amplify and sustain H2O2 production in large quantities, contributing to prolonged pathological signaling.

3 Hydrogen peroxide regulation of endothelial cell growth and proliferation

H2O2 generated by xanthine oxidase or glucose oxidase stimulated endothelial cell growth [41]. Consistently, scavenging H2O2 by adenoviral overexpression of catalase or cytosolic glutathione peroxidase inhibited endothelial cell proliferation [42,43]. These data indicate a role of H2O2 in growth signaling. Indeed, H2O2 is downstream of Flk1/KDR, proceeding activation of ERK1/2 [44]. The growth regulating p90RSK, an important downstream effector of ERK1/2, is activated by H2O2 in endothelial cells [45]. Furthermore, H2O2 feed-forwardly increases endothelial VEGF expression and secretion via activation of PKC and NFκB [46], and regulates expression and function of other growth factors [47]. Growth signaling induced by agonists other than growth factors may also require H2O2. For instance, cyclic strain induces early growth factor 1 (Egr-1) expression and this requires H2O2-dependent activation of ERK1/2 [48].

Furthermore, H2O2 increases activity of eIF4E, an eukaryotic initiation factor of protein synthesis [49], and is effective in promoting endothelial cell tube formation [50,51]. Hypoxia induced activation of ERK1/2 and p38 MAPK in lung endothelial cells is mediated by H2O2[52]. This may partially underlie hypoxia-induced angiogenesis. Indeed, endogenously produced H2O2 upregulates VEGF expression and angiogenesis in vivo [53]. The signaling events involved in H2O2 modulation of endothelial cell growth, and later discussed endothelial apoptosis are illustrated in Fig. 2. Whereas a modest increase in endothelial cell growth and angiogenesis is beneficial in repairing ischemic damage [54], excessive growth contributes to atheroma formation post injury [53]. In addition, when H2O2 diffuses to adjacent vascular smooth muscle, it could induce hypertrophy [13]. Indeed, endogenously overproduced H2O2 augmented angiotensin II induced vascular hypertrophy [55].

Signaling events mediating hydrogen peroxide modulation of endothelial cell growth and apoptosis. Mitogenic stimuli such as vascular endothelial growth factor (VEGF) and cyclic strain activate vascular NAD(P)H oxidases to form superoxide (O2.−) and subsequently hydrogen peroxide (H2O2). H2O2 in turn mediates activations of p38 MAPK, ERK1/2 and transcriptional factors involved in grow signaling. Of interest, H2O2 also feed-forwardly upregulates endothelial cell expression of VEGF. On the other hand, excessively produced H2O2 exceeding 200 μmol/L induces endothelial apoptosis. This response seems to involve transferrin receptor (TfR)-dependent intracellular iron uptake, and activations of mitochondrion, FAS and JNK/c-Jun pathway.
Fig. 2

Signaling events mediating hydrogen peroxide modulation of endothelial cell growth and apoptosis. Mitogenic stimuli such as vascular endothelial growth factor (VEGF) and cyclic strain activate vascular NAD(P)H oxidases to form superoxide (O2.−) and subsequently hydrogen peroxide (H2O2). H2O2 in turn mediates activations of p38 MAPK, ERK1/2 and transcriptional factors involved in grow signaling. Of interest, H2O2 also feed-forwardly upregulates endothelial cell expression of VEGF. On the other hand, excessively produced H2O2 exceeding 200 μmol/L induces endothelial apoptosis. This response seems to involve transferrin receptor (TfR)-dependent intracellular iron uptake, and activations of mitochondrion, FAS and JNK/c-Jun pathway.

4 Hydrogen peroxide regulation of endothelial apoptosis

Cultured umbilical vein endothelial cells are particularly prone to oxidative damage and often used to study H2O2 induced apoptosis. Of note, aortic endothelial cells undergo no significant cell death with exogenous addition of H2O2 less than 200 μmol/L [56–58]. However, 50–100 μmol/L H2O2 was apoptotic for umbilical vein endothelial cells. Endogenous H2O2, stimulated by oxidized LDL, induced endothelial cell apoptosis via JNK activation [59]. Mitochondrion-dependent ROS propagation is likely involved because mitochondria-targeted antioxidant abrogated H2O2 activation of JNK and apoptosis [60]. Studies by other groups also support a role for JNK/c-Jun in H2O2 induced apoptosis of endothelial cells [59,61,62]. Of note, Chen et al. previously characterized upstream events of JNK activation by H2O2, which involve Src-dependent activation of EGFR [63]. Additionally, H2O2 upregulates Fas expression [64], enhances intracellular ion uptake [65], and induces mitochondrial DNA damage [66], all of which may directly trigger apoptosis.

Importantly, endothelial cell apoptosis has been implicated in atherosclerosis [67,68] although molecular mechanisms underlying H2O2 induced endothelial cell apoptosis remain to be fully elucidated. It is however interesting to note that H2O2 induced apoptosis is inhibitable by unidirectional laminar shear stress, which exerts this effect by attenuating JNK activation and enhancing glutathione reductase-dependent glutathione redox-cycling [69,70]. Furthermore, S-nitrosylation of thioredoxin 1 [71] and ERK5 activation [72] was found protective of endothelial cells from apoptosis.

5 Hydrogen peroxide regulation of endothelial barrier function and actin cytoskeleton

Relatively high concentrations of H2O2 (>200 μmol/L) increase endothelial permeability [73], and this seems mediated by activations of PKC [74], phosphodiesterase [75], small G protein Rho [76], Src tyrosine kinase [77], p38 MAPK [78], and an increase in intracellular calcium [79]. Tight junction protein occludin was found downstream of ERK1/2 in mediating H2O2-induced barrier dysfunction [80]. In addition, H2O2 downregulation of intracellular cAMP content appears important in inducing barrier dysfunction [75,81]. Decreased PKA activity can itself account for many of the activating effects that result in barrier dysfunction or at least permit those changes [82]. Crosstalks or spatial regulations among above-described signaling events however remain to be revealed. Caveolin-1 is expressed abundantly in lipid rafts of endothelial cell membrane, where it plays an important role in modulating barrier function [83]. Whether caveolin-1 is required for H2O2 modulation of endothelial barrier function however is unclear although our unpublished data suggest that H2O2 induces rapid Src-dependent phosphorylation of caveolin-1. Of note, at various concentrations, NO. is able to enhance or offset H2O2-induced loss of endothelial barrier function [84,85].

It has been shown that cytoskeletal reorganization modulates vascular permeability [86]. Likewise, H2O2 regulation of barrier function is linked to its effects on endothelial actin cytoskeleton. p38 MAPK phosphorylation of actin binding protein HSP27 mediates reorganization of actin cytoskeleton in H2O2-stimulated endothelial cells [87]. In keeping with this, H2O2 induced formation of actin stress fibers was found to be mediated by CaMKII-dependent activation of p38 MAPK/HSP27 and ERK1/2 [88]. Whereas HSP27 phosphorylation was completely prevented by blockade of p38 MAPK, inhibition of ERK1/2 only transiently attenuated HSP27 phosphorylation [88]. These data suggest that ERK1/2 has different downstream effectors in modulating actin cytoskeleton.

It was previously shown that endothelial MLCK is involved in H2O2 induced actin reorganization [89]. MLCK also plays an important role in thrombin induced endothelial barrier dysfunction [86]. Thrombin has been shown to stimulate endothelial NAD(P)H oxidase activation and subsequent H2O2 production [90]. Recent studies further demonstrated that thrombin upregulates p22phox expression via p38 MAPK and PI-3K/AKT [91]. Given these observations, it is interesting to speculate that MLCK lies downstream of ERK1/2 in mediating H2O2 induced actin reorganization. In addition, H2O2 phosphorylation of tropomyosin-1 is ERK1/2-dependent, preceding its co-localization with developing actin filaments [92]. Of note, H2O2 activation of phospholipase D is also involved in actin reorganization in endothelial cells [93], likely serving as upstream activator of ERK1/2 [94]. The signaling events described above are summarized as part of the Fig. 3. Importantly, H2O2 regulation of endothelial cytoskeleton and barrier function may not only mediate ischemia–reperfusion induced acute vascular injury [73], but also potentiate vascular inflammation to contribute to chronic vascular disease [95].

Signaling events mediating hydrogen peroxide modulation of endothelial actin cytoskeleton and barrier function. Intracellular hydrogen peroxide (H2O2) produced in response to extracellular stimuli such as thrombin is capable of inducing phospholipase D (PLD) or calcium/calmodulin-dependent protein kinase II (CaMKII) dependent activation of ERK1/2, and CaMKII-dependent activation of p38 MAPK. By activating downstream effectors tropomyosin-1/MLCK or heat shock protein 27 (HSP27) respectively, both pathways are involved in H2O2 induced reorganization of actin cytoskeleton, which is associated with modulation of barrier function. Various kinases or signaling intermediates such as PKC, PKA and calcium have been implicated in H2O2 induction of endothelial barrier dysfunction. The inter-relationships among these players however remain to be elucidated.
Fig. 3

Signaling events mediating hydrogen peroxide modulation of endothelial actin cytoskeleton and barrier function. Intracellular hydrogen peroxide (H2O2) produced in response to extracellular stimuli such as thrombin is capable of inducing phospholipase D (PLD) or calcium/calmodulin-dependent protein kinase II (CaMKII) dependent activation of ERK1/2, and CaMKII-dependent activation of p38 MAPK. By activating downstream effectors tropomyosin-1/MLCK or heat shock protein 27 (HSP27) respectively, both pathways are involved in H2O2 induced reorganization of actin cytoskeleton, which is associated with modulation of barrier function. Various kinases or signaling intermediates such as PKC, PKA and calcium have been implicated in H2O2 induction of endothelial barrier dysfunction. The inter-relationships among these players however remain to be elucidated.

6 Hydrogen peroxide regulation of endothelium-dependent vasorelaxation and eNOS expression

Earlier studies demonstrated that H2O2 produces endothelium-dependent and independent vasorelaxation [96,97]. The endothelium-dependent vasorelaxation caused by H2O2 in large vessels depends on eNOS, as L-NAME (NOS inhibitor) abolishes it [98,99]. Indeed, H2O2 can acutely stimulate eNOS production of NO. via PI3-K/AKT and ERK1/2, [58,100]. When eNOS was uncoupled to produce O2.− rather than NO. in hypertensive or atherosclerotic large vessels, endothelium-derived H2O2 mediated compensatory relaxation via unknown mechanisms [10,101]. One possibility is direct polarization of vascular smooth muscle.

Indeed, H2O2 was found to be an endothelium-derived hyperpolarizing factor in small arteries [102,103] and an activator of potassium channel in large cerebral arteries [104]. In human coronary arterioles, flow-induced vasodilatation is mediated by endothelium-derived H2O2[105]. In these arterioles, the enzymatic source of H2O2 appears to be the mitochondrial respiratory chain whereas in large vessels, vascular NAD(P)H oxidases are responsible for H2O2 production [6,8]. This seems consistent with previous observations that NO. plays a lesser role in vasodilatation of small arteries. Of note, in coronary arterioles, H2O2 induced endothelium-dependent vasorelaxation is NO.-independent involving activation of cyclooxygenase 1 and smooth muscle potassium channel [106]. Therefore, it is clear that H2O2 is capable of mediating endothelium-dependent vasorelaxation, NO.-dependently or independently. The underlying mechanisms however are diverse depending on the size of the blood vessels and the availability of functional eNOS to produce NO.[107,108].

H2O2 can also stimulate endothelium-independent vasorelaxation with unknown mechanisms [109–111]. Early studies demonstrated that H2O2 can directly activate cyclic GMP via a compound I/H2O2 complex [109,110]. Additional recent studies suggest that this might be true [112]. Known and hypothetical mechanisms underlying H2O2 induced vasorelaxation are summarized schematically in Fig. 4.

Mechanisms underlying endothelium-dependent or independent vasorelaxation induced by hydrogen peroxide. Depending on size of the blood vessels and availability of bioactive nitric oxide (NO.), and physiological versus pathological environment, hydrogen peroxide (H2O2) mediates endothelium-dependent or independent vasorelaxation via NO.-dependent or independent mechanisms. H2O2 seems to activate eNOS under physiological conditions in large vessels, resulting in endothelium and NO.-dependent relaxation. In small vessels such as coronary arterioles, mitochondrial respiratory chain-derived H2O2 is found to be responsible for flow-mediated vasodilatation that is independent of NO.. Under pathological conditions such as atherosclerosis and hypertension, H2O2 produced by large vessels mediates compensatory, endothelium-dependent but NO.-independent relaxation. Additionally, under unclear conditions, H2O2 may also cause endothelium-independent relaxation via compound I-mediated direct activation of smooth muscle cyclic GMP.
Fig. 4

Mechanisms underlying endothelium-dependent or independent vasorelaxation induced by hydrogen peroxide. Depending on size of the blood vessels and availability of bioactive nitric oxide (NO.), and physiological versus pathological environment, hydrogen peroxide (H2O2) mediates endothelium-dependent or independent vasorelaxation via NO.-dependent or independent mechanisms. H2O2 seems to activate eNOS under physiological conditions in large vessels, resulting in endothelium and NO.-dependent relaxation. In small vessels such as coronary arterioles, mitochondrial respiratory chain-derived H2O2 is found to be responsible for flow-mediated vasodilatation that is independent of NO.. Under pathological conditions such as atherosclerosis and hypertension, H2O2 produced by large vessels mediates compensatory, endothelium-dependent but NO.-independent relaxation. Additionally, under unclear conditions, H2O2 may also cause endothelium-independent relaxation via compound I-mediated direct activation of smooth muscle cyclic GMP.

Of note, H2O2 potently upregulates eNOS expression in vitro and in vivo [37,56,113]. Oscillatory shear stress upregulation of eNOS requires H2O2 activation of CaMKII [114]. These observations are consistent with previous findings that eNOS is upregulated in animal models of atherosclerosis, diabetes or aging, where vascular H2O2 production is increased [115,116]. Rapidly accumulating evidence suggests that eNOS can become uncoupled to produce O2.− under conditions such as hypertension and atherosclerosis, likely consequent to tetrahydrobiopterin deficiency and oxidation. Thus upregulation of eNOS could be less beneficial or even detrimental while functioning as an oxidase. Indeed, our recent study indicates that endothelial NAD(P)H oxidase-derived H2O2 downregulates tetrahydrobiopterin salvage enzyme dihydrofolate reductase (DHFR) and DHFR/eNOS ratio in response to angiotensin II, leading to tetrahydrobiopterin deficiency and subsequently uncoupling of eNOS [9].

7 Hydrogen peroxide regulation of endothelial inflammatory responses

Exposure of endothelial cells to H2O2 (50–100 μmol/L) increased surface expression of intercellular adhesion molecule-1 (ICAM-1) [117]. Furthermore, endogenously produced H2O2 mediates ischemia–reperfusion induced upregulation of ICAM-1 [118] and MCP-1 (monocyte chemoattractant protein-1) [119]. H2O2 is also required for TNFα induction of ICAM-1 and VCAM-1 (vascular cell adhesion molecule-1) [120], and upregulation of MCP-1 by TNFα [121] or hyperglycemia [122]. Additionally, activation of NFκB by H2O2 was found responsible for TNFα induction of ICAM-1 [123] and angiotensin II upregulation of VCAM-1 [124].

Endothelial expression of platelet activating factor (PAF) [125] and P-selectin [126] is inducible by H2O2. All these molecules including ICAM-1, VCAM-1, MCP-1, PAF and P-selectin have been shown to mediate neutrophil adhesion to endothelium (Fig. 5). PAF and P-selectin also mediates platelet activation and endothelium–platelet interaction. Upregulation of Mac 1 (CD11b and CD18) in neutrophils [127] by H2O2 stimulates more H2O2 release from activated neutrophils [128], forming a vicious circle. Taken together, these data strongly suggest that H2O2 mediates endothelial expression of inflammatory proteins and enhanced platelet–endothelium interaction, which in turn contributes to development of atherosclerotic vascular disease. The events described above are summarized in Fig. 5.

Mechanisms whereby hydrogen peroxide promotes endothelial inflammatory responses. By activating redox-sensitive transcription factors such as NFκB, hydrogen peroxide (H2O2) mediates endothelial induction of inflammatory proteins such as VCAM, ICAM and MCP-1 in response to extracellular stimuli such as TNFα, hypoxia and angiotensin II. H2O2 also activates platelet and neutrophil interactions with endothelium to facilitate inflammation of the endothelium.
Fig. 5

Mechanisms whereby hydrogen peroxide promotes endothelial inflammatory responses. By activating redox-sensitive transcription factors such as NFκB, hydrogen peroxide (H2O2) mediates endothelial induction of inflammatory proteins such as VCAM, ICAM and MCP-1 in response to extracellular stimuli such as TNFα, hypoxia and angiotensin II. H2O2 also activates platelet and neutrophil interactions with endothelium to facilitate inflammation of the endothelium.

8 Hydrogen peroxide regulation of endothelium-mediated vascular remodeling

Emerging evidence seems to suggest that H2O2 plays an important role in mediating vascular remodeling. First of all, H2O2 increases endothelial MMP-2 expression [12]. TGF-β1 and macrophages are important regulators of tissue fibrosis and remodeling. TGF-β1 induction of macrophage colony-stimulating factor, and IL-1 induction of plasminogen activator inhibitor-1 (PAI-1) in endothelial cells were found dependent on H2O2 activation of NFκB [129,130]. PAI-1 plays an important role in vascular remodeling induced by thrombosis [131] or chronic eNOS inhibition [132]. Recent studies demonstrated that PAI-1 is also critically important for angiotensin II or salt-induced vascular remodeling in vivo [133]. Thus by mediating TGF-β1 and PAI-1 signaling and MMP release, H2O2 contributes to endothelium-regulated vascular remodeling, which has been shown to regulate stability of atherosclerotic plaques [134].

9 Contribution of hydrogen peroxide to vascular disease

To date, the specific, individual ROS that is most relevant to vascular signaling pathophysiologically is yet identified. Nevertheless, selectively overproducing or removing H2O2 significantly altered atherogenesis in animal models. Mice overexpressing p22phox had markedly increased atheroma formation in a carotid ligation model [53]. This response was associated with enhanced H2O2 production in the vessel wall, and was abolished by scavenging H2O2 with ebselen, implicating a critical role of H2O2[53]. Parallel studies from another group confirmed the same notion [135]. Yang et al. cross bred transgenic mice overexpressing Cu/Zn-SOD or catalase, with mice deficient in apolipoprotein E (apoE−/−), to examine a specific role of H2O2 versus O2.− in atherogenesis [135]. They found that whereas overexpressing Cu/Zn-SOD had no effect on atherosclerotic lesion formation, overexpression of catalase, or co-overexpression of catalase and Cu/Zn-SOD markedly retarded atherosclerosis in all aspects [135]. These observations were consistent with the findings by Tribble et al. that overexpression of Cu/Zn-SOD failed to prevent atherosclerosis in high fat diet-fed apoE−/− mice [136]. Taken together, these data suggest that H2O2 is more atherogenic than O2.−. It is true that Cu/Zn-SOD is intracellular and overexpression of Cu/Zn-SOD can not protect NO. during its trafficking to vascular smooth muscle. Thus extracellular SOD might be more important in determining NO. bioavailability and be more anti-atherosclerosis. Indeed, evidence gained from ecSOD-null mice and adenovirus-mediated overexpression of ecSOD supports that ecSOD is the main determinant of NO. bioavailability in the vessel wall, and is thus involved in blood pressure regulation [137,138]. However, impact of ecSOD overexpression on atherosclerosis is yet to be reported although mice deficient in ecSOD developed similar atherosclerotic lesions compared to wild-type mice [139]. Therefore, whereas O2.− is important in directly modulating NO. bioavailability and serving as the precursor for H2O2[7], relatively lasting H2O2 seems more important in mediating atherogenic signaling.

In summary, O2.− and H2O2 are produced in vascular cells by multiple enzymatic systems including vascular NAD(P)H oxidases, mitochondrion, xanthine oxidase and uncoupled eNOS. While some O2.− spontaneously degrades by reacting with NO., O2.− signal preserved by dismutation into H2O2 exerts prolonged signaling effects. This may explain why direct scavenging H2O2 but not O2.− is more effective in athero-protection. Although there is much to be learned, the impact of H2O2 on different aspects of endothelial cell function as discussed throughout the review may at least partially underlie H2O2-mediated development of atherosclerotic vascular disease.

Acknowledgements

The author's work is supported by an American Heart Association Scientist Development Grant (#0435189N), an American Diabetes Association Research Award, a Career Development Award from the Schweppe Foundation, and a Start-up Fund from the University of Chicago.

References

[1]

Cai
H.

NAD(P)H oxidase-dependent self-propagation of hydrogen peroxide and vascular disease
Circ Res
2005
96
818
822
[2]

Brandes
R.P.
Kreuzer
J.

Vascular NADPH oxidases: molecular mechanisms of activation
Cardiovasc Res
2005
65
16
27
[3]

Li
J.M.
Shah
A.M.

Endothelial cell superoxide generation: regulation and relevance for cardiovascular pathophysiology
Am J Physiol Regul Integr Comp Physiol
2004
287
R1014
R1030
[4]

Griendling
K.K.

Novel NAD(P)H oxidases in the cardiovascular system
Heart
2004
90
491
493
[5]

Touyz
R.M.
Schiffrin
E.L.

Reactive oxygen species in vascular biology: implications in hypertension
Histochem Cell Biol
2004
122
339
352
[6]

Cai
H.
Griendling
K.K.
Harrison
D.G.

The vascular NAD(P)H oxidases as therapeutic targets in cardiovascular diseases
Trends Pharmacol Sci
2003
24
471
478
[7]

Cai
H.
Harrison
D.G.

Endothelial dysfunction in cardiovascular diseases: the role of oxidant stress
Circ Res
2000
87
840
844
[8]

Liu
Y.
Zhao
H.
Li
H.
Kalyanaraman
B.
Nicolosi
A.C.
Gutterman
D.D.

Mitochondrial sources of H2O2 generation play a key role in flow-mediated dilation in human coronary resistance arteries
Circ Res
2003
93
573
580
[9]

Chalupsky
K.
Cai
H.

Endothelial dihydrofolate reductase: critical for nitric oxide bioavailability and role in angiotensin II uncoupling of eNOS
Proc Natl Acad Sci U S A
2005
102
9056
9061
[10]

Landmesser
U.
Dikalov
S.
Price
S.R.
McCann
L.
Fukai
T.
Holland
S.M.
et al. 

Oxidation of tetrahydrobiopterin leads to uncoupling of endothelial cell nitric oxide synthase in hypertension
J Clin Invest
2003
111
1201
1209
[11]

McNally
J.S.
Davis
M.E.
Giddens
D.P.
Saha
A.
Hwang
J.
Dikalov
S.
et al. 

Role of xanthine oxidoreductase and NAD(P)H oxidase in endothelial superoxide production in response to oscillatory shear stress
Am J Physiol Heart Circ Physiol
2003
285
H2290
H2297
[12]

Belkhiri
A.
Richards
C.
Whaley
M.
McQueen
S.A.
Orr
F.W.

Increased expression of activated matrix metalloproteinase-2 by human endothelial cells after sublethal H2O2 exposure
Lab Invest
1997
77
533
539
[13]

Zafari
A.M.
Ushio-Fukai
M.
Akers
M.
Yin
Q.
Shah
A.
Harrison
D.G.
et al. 

Role of NADH/NADPH oxidase-derived H2O2 in angiotensin II-induced vascular hypertrophy
Hypertension
1998
32
488
495
[14]

Brennan
M.L.
Wu
W.
Fu
X.
Shen
Z.
Song
W.
Frost
H.
et al. 

A tale of two controversies: defining both the role of peroxidases in nitrotyrosine formation in vivo using eosinophil peroxidase and myeloperoxidase-deficient mice, and the nature of peroxidase-generated reactive nitrogen species
J Biol Chem
2002
277
17415
17427
[15]

Baldus
S.
Eiserich
J.P.
Brennan
M.L.
Jackson
R.M.
Alexander
C.B.
Freeman
B.A.

Spatial mapping of pulmonary and vascular nitrotyrosine reveals the pivotal role of myeloperoxidase as a catalyst for tyrosine nitration in inflammatory diseases
Free Radic Biol Med
2002
33
1010
[16]

Bayraktutan
U.
Blayney
L.
Shah
A.M.

Molecular characterization and localization of the NAD(P)H oxidase components gp91-phox and p22-phox in endothelial cells
Arterioscler Thromb Vasc Biol
2000
20
1903
1911
[17]

Li
J.M.
Shah
A.M.

Intracellular localization and preassembly of the NADPH oxidase complex in cultured endothelial cells
J Biol Chem
2002
277
19952
19960
[18]

Frey
R.S.
Rahman
A.
Kefer
J.C.
Minshall
R.D.
Malik
A.B.

PKCzeta regulates TNF-alpha-induced activation of NADPH oxidase in endothelial cells
Circ Res
2002
90
1012
1019
[19]

Li
J.M.
Shah
A.M.

Mechanism of endothelial cell NADPH oxidase activation by angiotensin II: role of the p47phox subunit
J Biol Chem
2003
278
12094
12100
[20]

Gorlach
A.
Brandes
R.P.
Nguyen
K.
Amidi
M.
Dehghani
F.
Busse
R.

A gp91phox containing NADPH oxidase selectively expressed in endothelial cells is a major source of oxygen radical generation in the arterial wall
Circ Res
2000
87
26
32
[21]

Furst
R.
Brueckl
C.
Kuebler
W.M.
Zahler
S.
Krotz
F.
Gorlach
A.
et al. 

Atrial natriuretic peptide induces mitogen-activated protein kinase phosphatase-1 in human endothelial cells via Rac1 and NAD(P)H oxidase/Nox2-activation
Circ Res
2005
96
43
53
[22]

Ago
T.
Kitazono
T.
Ooboshi
H.
Iyama
T.
Han
Y.H.
Takada
J.
et al. 

Nox4 as the major catalytic component of an endothelial NAD(P)H oxidase
Circulation
2004
109
227
233
[23]

Buul
J.D.
Fernandez-Borja
M.
Anthony
E.C.
Hordijk
P.L.

Expression and localization of NOX2 and NOX4 in primary human endothelial cells
Antioxid Redox Signal
2005
7
308
317
[24]

Sorescu
G.P.
Song
H.
Tressel
S.L.
Hwang
J.
Dikalov
S.
Smith
D.A.
et al. 

Bone morphogenic protein 4 produced in endothelial cells by oscillatory shear stress induces monocyte adhesion by stimulating reactive oxygen species production from a nox1-based NADPH oxidase
Circ Res
2004
95
773
779
[25]

Kobayashi
S.
Nojima
Y.
Shibuya
M.
Maru
Y.

Nox1 regulates apoptosis and potentially stimulates branching morphogenesis in sinusoidal endothelial cells
Exp Cell Res
2004
300
455
462
[26]

Lassegue
B.
Sorescu
D.
Szocs
K.
Yin
Q.
Akers
M.
Zhang
Y.
et al. 

Novel gp91(phox) homologues in vascular smooth muscle cells: nox1 mediates angiotensin II-induced superoxide formation and redox-sensitive signaling pathways
Circ Res
2001
88
888
894
[27]

Geiszt
M.
Kopp
J.B.
Varnai
P.
Leto
T.L.

Identification of renox, an NAD(P)H oxidase in kidney
Proc Natl Acad Sci U S A
2000
97
8010
8014
[28]

Hilenski
L.L.
Clempus
R.E.
Quinn
M.T.
Lambeth
J.D.
Griendling
K.K.

Distinct subcellular localizations of Nox1 and Nox4 in vascular smooth muscle cells
Arterioscler Thromb Vasc Biol
2004
24
677
683
[29]

Geiszt
M.
Lekstrom
K.
Witta
J.
Leto
T.L.

Proteins homologous to p47phox and p67phox support superoxide production by NAD(P)H oxidase 1 in colon epithelial cells
J Biol Chem
2003
278
20006
20012
[30]

Banfi
B.
Clark
R.A.
Steger
K.
Krause
K.H.

Two novel proteins activate superoxide generation by the NADPH oxidase NOX1
J Biol Chem
2003
278
3510
3513
[31]

Takeya
R.
Ueno
N.
Kami
K.
Taura
M.
Kohjima
M.
Izaki
T.
et al. 

Novel human homologues of p47phox and p67phox participate in activation of superoxide-producing NADPH oxidases
J Biol Chem
2003
278
25234
25246
[32]

Kawahara
T.
Kuwano
Y.
Teshima-Kondo
S.
Takeya
R.
Sumimoto
H.
Kishi
K.
et al. 

Role of nicotinamide adenine dinucleotide phosphate oxidase 1 in oxidative burst response to toll-like receptor 5 signaling in large intestinal epithelial cells
J Immunol
2004
172
3051
3058
[33]

Edens
W.A.
Sharling
L.
Cheng
G.
Shapira
R.
Kinkade
J.M.
Lee
T.
et al. 

Tyrosine cross-linking of extracellular matrix is catalyzed by Duox, a multidomain oxidase/peroxidase with homology to the phagocyte oxidase subunit gp91phox
J Cell Biol
2001
154
879
891
[34]

Lambeth
J.D.

Nox/Duox family of nicotinamide adenine dinucleotide (phosphate) oxidases
Curr Opin Hematol
2002
9
11
17
[35]

Geiszt
M.
Witta
J.
Baffi
J.
Lekstrom
K.
Leto
T.L.

Dual oxidases represent novel hydrogen peroxide sources supporting mucosal surface host defense
FASEB J
2003
17
1502
1504
[36]

Forteza
R.
Salathe
M.
Miot
F.
Conner
G.E.

Regulated H2O2 production by Duox in human airway epithelial cells
Am J Respir Cell Mol Biol
2005
32
462
469
[37]

Laude
K.
Cai
H.
Fink
B.
Hoch
N.
Weber
D.S.
McCann
L.
et al. 

Hemodynamic and biochemical adaptations to vascular smooth muscle overexpression of p22phox in mice
Am J Physiol Heart Circ Physiol
2005
288
H7
H12
[38]

Hanna
I.R.
Hilenski
L.L.
Dikalova
A.
Taniyama
Y.
Dikalov
S.
Lyle
A.
et al. 

Functional association of nox1 with p22phox in vascular smooth muscle cells
Free Radic Biol Med
2004
37
1542
1549
[39]

Ambasta
R.K.
Kumar
P.
Griendling
K.K.
Schmidt
H.H.
Busse
R.
Brandes
R.P.

Direct interaction of the novel Nox proteins with p22phox is required for the formation of a functionally active NADPH oxidase
J Biol Chem
2004
279
45935
45941
[40]

Guzik
T.J.
Sadowski
J.
Kapelak
B.
Jopek
A.
Rudzinski
P.
Pillai
R.
et al. 

Systemic regulation of vascular NAD(P)H oxidase activity and nox isoform expression in human arteries and veins
Arterioscler Thromb Vasc Biol
2004
24
1614
1620
[41]

Ruiz-Gines
J.A.
Lopez-Ongil
S.
Gonzalez-Rubio
M.
Gonzalez-Santiago
L.
Rodriguez-Puyol
M.
Rodriguez-Puyol
D.

Reactive oxygen species induce proliferation of bovine aortic endothelial cells
J Cardiovasc Pharmacol
2000
35
109
113
[42]

Zanetti
M.
Katusic
Z.S.
O'Brien
T.

Adenoviral-mediated overexpression of catalase inhibits endothelial cell proliferation
Am J Physiol Heart Circ Physiol
2002
283
H2620
H2626
[43]

Faucher
K.
Rabinovitch-Chable
H.
Barriere
G.
Cook-Moreau
J.
Rigaud
M.

Overexpression of cytosolic glutathione peroxidase (GPX1) delays endothelial cell growth and increases resistance to toxic challenges
Biochimie
2003
85
611
617
[44]

Colavitti
R.
Pani
G.
Bedogni
B.
Anzevino
R.
Borrello
S.
Waltenberger
J.
et al. 

Reactive oxygen species as downstream mediators of angiogenic signaling by vascular endothelial growth factor receptor-2/KDR
J Biol Chem
2002
277
3101
3108
[45]

Abe
J.
Okuda
M.
Huang
Q.
Yoshizumi
M.
Berk
B.C.

Reactive oxygen species activate p90 ribosomal S6 kinase via Fyn and Ras
J Biol Chem
2000
275
1739
1748
[46]

Chua
C.C.
Hamdy
R.C.
Chua
B.H.

Upregulation of vascular endothelial growth factor by H2O2 in rat heart endothelial cells
Free Radic Biol Med
1998
25
891
897
[47]

Eyries
M.
Collins
T.
Khachigian
L.M.

Modulation of growth factor gene expression in vascular cells by oxidative stress
Endothelium
2004
11
133
139
[48]

Wung
B.S.
Cheng
J.J.
Chao
Y.J.
Hsieh
H.J.
Wang
D.L.

Modulation of Ras/Raf/extracellular signal-regulated kinase pathway by reactive oxygen species is involved in cyclic strain-induced early growth response-1 gene expression in endothelial cells
Circ Res
1999
84
804
812
[49]

Duncan
R.F.
Peterson
H.
Hagedorn
C.H.
Sevanian
A.

Oxidative stress increases eukaryotic initiation factor 4E phosphorylation in vascular cells
Biochem J
2003
369
213
225
[50]

Shono
T.
Ono
M.
Izumi
H.
Jimi
S.I.
Matsushima
K.
Okamoto
T.
et al. 

Involvement of the transcription factor NF-kappaB in tubular morphogenesis of human microvascular endothelial cells by oxidative stress
Mol Cell Biol
1996
16
4231
4239
[51]

Yasuda
M.
Ohzeki
Y.
Shimizu
S.
Naito
S.
Ohtsuru
A.
Yamamoto
T.
et al. 

Stimulation of in vitro angiogenesis by hydrogen peroxide and the relation with ETS-1 in endothelial cells
Life Sci
1999
64
249
258
[52]

Parinandi
N.L.
Kleinberg
M.A.
Usatyuk
P.V.
Cummings
R.J.
Pennathur
A.
Cardounel
A.J.
et al. 

Hyperoxia-induced NAD(P)H oxidase activation and regulation by MAP kinases in human lung endothelial cells
Am J Physiol Lung Cell Mol Physiol
2003
284
L26
L38
[53]

Khatri
J.J.
Johnson
C.
Magid
R.
Lessner
S.M.
Laude
K.M.
Dikalov
S.I.
et al. 

Vascular oxidant stress enhances progression and angiogenesis of experimental atheroma
Circulation
2004
109
520
525
[54]

Sellke
F.W.
Simons
M.

Angiogenesis in cardiovascular disease: current status and therapeutic potential
Drugs
1999
58
391
396
[55]

Weber
D.S.
Rocic
P.
Mellis
A.M.
Laude
K.
Lyle
A.N.
Harrison
D.G.
et al. 

Angiotensin II-induced hypertrophy is potentiated in mice overexpressing p22phox in vascular smooth muscle
Am J Physiol Heart Circ Physiol
2005
288
H37
H42
[56]

Cai
H.
Davis
M.E.
Drummond
G.R.
Harrison
D.G.

Induction of endothelial NO synthase by hydrogen peroxide via a Ca(2+)/calmodulin-dependent protein kinase II/janus kinase 2-dependent pathway
Arterioscler Thromb Vasc Biol
2001
21
1571
1576
[57]

Cai
H.
Li
Z.
Dikalov
S.
Holland
S.M.
Hwang
J.
Jo
H.
et al. 

NAD(P)H oxidase-derived hydrogen peroxide mediates endothelial nitric oxide production in response to angiotensin II
J Biol Chem
2002
277
48311
48317
[58]

Thomas
S.R.
Chen
K.
Keaney
J.F.
Jr.

Hydrogen peroxide activates endothelial nitric-oxide synthase through coordinated phosphorylation and dephosphorylation via a phosphoinositide 3-kinase-dependent signaling pathway
J Biol Chem
2002
277
6017
6024
[59]

Lin
S.J.
Shyue
S.K.
Liu
P.L.
Chen
Y.H.
Ku
H.H.
Chen
J.W.
et al. 

Adenovirus-mediated overexpression of catalase attenuates oxLDL-induced apoptosis in human aortic endothelial cells via AP-1 and C-Jun N-terminal kinase/extracellular signal-regulated kinase mitogen-activated protein kinase pathways
J Mol Cell Cardiol
2004
36
129
139
[60]

Chen
K.
Thomas
S.R.
Albano
A.
Murphy
M.P.
Keaney
J.F.
Jr.

Mitochondrial function is required for hydrogen peroxide-induced growth factor receptor transactivation and downstream signaling
J Biol Chem
2004
279
35079
35086
[61]

Wang
N.
Verna
L.
Hardy
S.
Zhu
Y.
Ma
K.S.
Birrer
M.J.
et al. 

c-Jun triggers apoptosis in human vascular endothelial cells
Circ Res
1999
85
387
393
[62]

Ramachandran
A.
Moellering
D.
Go
Y.M.
Shiva
S.
Levonen
A.L.
Jo
H.
et al. 

Activation of c-Jun N-terminal kinase and apoptosis in endothelial cells mediated by endogenous generation of hydrogen peroxide
Biol Chem
2002
383
693
701
[63]

Chen
K.
Vita
J.A.
Berk
B.C.
Keaney
J.F.
Jr.

c-Jun N-terminal kinase activation by hydrogen peroxide in endothelial cells involves SRC-dependent epidermal growth factor receptor transactivation
J Biol Chem
2001
276
16045
16050
[64]

Suhara
T.
Fukuo
K.
Sugimoto
T.
Morimoto
S.
Nakahashi
T.
Hata
S.
et al. 

Hydrogen peroxide induces up-regulation of Fas in human endothelial cells
J Immunol
1998
160
4042
4047
[65]

Kotamraju
S.
Tampo
Y.
Keszler
A.
Chitambar
C.R.
Joseph
J.
Haas
A.L.
et al. 

Nitric oxide inhibits H2O2-induced transferrin receptor-dependent apoptosis in endothelial cells: role of ubiquitin-proteasome pathway
Proc Natl Acad Sci U S A
2003
100
10653
10658
[66]

Ballinger
S.W.
Patterson
C.
Yan
C.N.
Doan
R.
Burow
D.L.
Young
C.G.
et al. 

Hydrogen peroxide- and peroxynitrite-induced mitochondrial DNA damage and dysfunction in vascular endothelial and smooth muscle cells
Circ Res
2000
86
960
966
[67]

Dimmeler
S.
Zeiher
A.M.

Vascular repair by circulating endothelial progenitor cells: the missing link in atherosclerosis?
J Mol Med
2004
82
671
677
[68]

Stoneman
V.E.
Bennett
M.R.

Role of apoptosis in atherosclerosis and its therapeutic implications
Clin Sci (Lond)
2004
107
343
354
[69]

Hermann
C.
Zeiher
A.M.
Dimmeler
S.

Shear stress inhibits H2O2-induced apoptosis of human endothelial cells by modulation of the glutathione redox cycle and nitric oxide synthase
Arterioscler Thromb Vasc Biol
1997
17
3588
3592
[70]

Hojo
Y.
Saito
Y.
Tanimoto
T.
Hoefen
R.J.
Baines
C.P.
Yamamoto
K.
et al. 

Fluid shear stress attenuates hydrogen peroxide-induced c-Jun NH2-terminal kinase activation via a glutathione reductase-mediated mechanism
Circ Res
2002
91
712
718
[71]

Haendeler
J.
Hoffmann
J.
Tischler
V.
Berk
B.C.
Zeiher
A.M.
Dimmeler
S.

Redox regulatory and anti-apoptotic functions of thioredoxin depend on S-nitrosylation at cysteine 69
Nat Cell Biol
2002
4
743
749
[72]

Pi
X.
Yan
C.
Berk
B.C.

Big mitogen-activated protein kinase (BMK1)/ERK5 protects endothelial cells from apoptosis
Circ Res
2004
94
362
369
[73]

Lum
H.
Barr
D.A.
Shaffer
J.R.
Gordon
R.J.
Ezrin
A.M.
Malik
A.B.

Reoxygenation of endothelial cells increases permeability by oxidant-dependent mechanisms
Circ Res
1992
70
991
998
[74]

Siflinger-Birnboim
A.
Goligorsky
M.S.
Del Vecchio
P.J.
Malik
A.B.

Activation of protein kinase C pathway contributes to hydrogen peroxide-induced increase in endothelial permeability
Lab Invest
1992
67
24
30
[75]

Suttorp
N.
Weber
U.
Welsch
T.
Schudt
C.

Role of phosphodiesterases in the regulation of endothelial permeability in vitro
J Clin Invest
1993
91
1421
1428
[76]

Hippenstiel
S.
Tannert-Otto
S.
Vollrath
N.
Krull
M.
Just
I.
Aktories
K.
et al. 

Glucosylation of small GTP-binding Rho proteins disrupts endothelial barrier function
Am J Physiol
1997
272
L38
L43
[77]

Kevil
C.G.
Okayama
N.
Alexander
J.S.

H(2)O(2)-mediated permeability II: importance of tyrosine phosphatase and kinase activity
Am J Physiol Cell Physiol
2001
281
C1940
C1947
[78]

Usatyuk
P.V.
Vepa
S.
Watkins
T.
He
D.
Parinandi
N.L.
Natarajan
V.

Redox regulation of reactive oxygen species-induced p38 MAP kinase activation and barrier dysfunction in lung microvascular endothelial cells
Antioxid Redox Signal
2003
5
723
730
[79]

Siflinger-Birnboim
A.
Lum
H.
Del Vecchio
P.J.
Malik
A.B.

Involvement of Ca2+ in the H2O2-induced increase in endothelial permeability
Am J Physiol
1996
270
L973
L978
[80]

Kevil
C.G.
Oshima
T.
Alexander
B.
Coe
L.L.
Alexander
J.S.

H(2)O(2)-mediated permeability: role of MAPK and occludin
Am J Physiol Cell Physiol
2000
279
C21
C30
[81]

Hastie
L.E.
Patton
W.F.
Hechtman
H.B.
Shepro
D.

H2O2-induced filamin redistribution in endothelial cells is modulated by the cyclic AMP-dependent protein kinase pathway
J Cell Physiol
1997
172
373
381
[82]

Patterson
C.E.
Lum
H.

Update on pulmonary edema: the role and regulation of endothelial barrier function
Endothelium
2001
8
75
105
[83]

Minshall
R.D.
Tiruppathi
C.
Vogel
S.M.
Malik
A.B.

Vesicle formation and trafficking in endothelial cells and regulation of endothelial barrier function
Histochem Cell Biol
2002
117
105
112
[84]

Chang
J.
Rao
N.V.
Markewitz
B.A.
Hoidal
J.R.
Michael
J.R.

Nitric oxide donor prevents hydrogen peroxide-mediated endothelial cell injury
Am J Physiol
1996
270
L931
L940
[85]

Okayama
N.
Kevil
C.G.
Correia
L.
Jourd'heuil
D.
Itoh
M.
Grisham
M.B.
et al. 

Nitric oxide enhances hydrogen peroxide-mediated endothelial permeability in vitro
Am J Physiol
1997
273
C1581
C1587
[86]

Dudek
S.M.
Garcia
J.G.

Cytoskeletal regulation of pulmonary vascular permeability
J Appl Physiol
2001
91
1487
1500
[87]

Huot
J.
Houle
F.
Marceau
F.
Landry
J.

Oxidative stress-induced actin reorganization mediated by the p38 mitogen-activated protein kinase/heat shock protein 27 pathway in vascular endothelial cells
Circ Res
1997
80
383
392
[88]

Nguyen
A.
Chen
P.
Cai
H.

Role of CaMKII in hydrogen peroxide activation of ERK1/2, p38 MAPK, HSP27 and actin reorganization in endothelial cells
FEBS Lett
2004
572
307
313
[89]

Zhao
Y.
Davis
H.W.

Hydrogen peroxide-induced cytoskeletal rearrangement in cultured pulmonary endothelial cells
J Cell Physiol
1998
174
370
379
[90]

Holland
J.A.
Meyer
J.W.
Chang
M.M.
O'Donnell
R.W.
Johnson
D.K.
Ziegler
L.M.

Thrombin stimulated reactive oxygen species production in cultured human endothelial cells
Endothelium
1998
6
113
121
[91]

Djordjevic
T.
Pogrebniak
A.
BelAiba
R.S.
Bonello
S.
Wotzlaw
C.
Acker
H.
et al. 

The expression of the NADPH oxidase subunit p22phox is regulated by a redox-sensitive pathway in endothelial cells
Free Radic Biol Med
2005
38
616
630
[92]

Houle
F.
Rousseau
S.
Morrice
N.
Luc
M.
Mongrain
S.
Turner
C.E.
et al. 

Extracellular signal-regulated kinase mediates phosphorylation of tropomyosin-1 to promote cytoskeleton remodeling in response to oxidative stress: impact on membrane blebbing
Mol Biol Cell
2003
14
1418
1432
[93]

Hastie
L.E.
Patton
W.F.
Hechtman
H.B.
Shepro
D.

Metabolites of the phospholipase D pathway regulate H2O2-induced filamin redistribution in endothelial cells
J Cell Biochem
1998
68
511
524
[94]

Andresen
B.T.
Rizzo
M.A.
Shome
K.
Romero
G.

The role of phosphatidic acid in the regulation of the Ras/MEK/Erk signaling cascade
FEBS Lett
2002
531
65
68
[95]

Siflinger-Birnboim
A.
Malik
A.B.

Regulation of endothelial permeability by second messengers
New Horiz
1996
4
87
98
[96]

Thomas
G.
Ramwell
P.

Induction of vascular relaxation by hydroperoxides
Biochem Biophys Res Commun
1986
139
102
108
[97]

Rubanyi
G.M.
Vanhoutte
P.M.

Oxygen-derived free radicals, endothelium, and responsiveness of vascular smooth muscle
Am J Physiol
1986
250
H815
H821
[98]

Zembowicz
A.
Hatchett
R.J.
Jakubowski
A.M.
Gryglewski
R.J.

Involvement of nitric oxide in the endothelium-dependent relaxation induced by hydrogen peroxide in the rabbit aorta
Br J Pharmacol
1993
110
151
158
[99]

Yang
Z.
Zhang
A.
Altura
B.T.
Altura
B.M.

Hydrogen peroxide-induced endothelium-dependent relaxation of rat aorta involvement of Ca2+ and other cellular metabolites
Gen Pharmacol
1999
33
325
336
[100]

Cai
H.
Li
Z.
Davis
M.E.
Kanner
W.
Harrison
D.G.
Dudley
S.C.
Jr.

Akt-dependent phosphorylation of serine 1179 and mitogen-activated protein kinase kinase/extracellular signal-regulated kinase 1/2 cooperatively mediate activation of the endothelial nitric-oxide synthase by hydrogen peroxide
Mol Pharmacol
2003
63
325
331
[101]

Laursen
J.B.
Somers
M.
Kurz
S.
McCann
L.
Warnholtz
A.
Freeman
B.A.
et al. 

Endothelial regulation of vasomotion in apoE-deficient mice: implications for interactions between peroxynitrite and tetrahydrobiopterin
Circulation
2001
103
1282
1288
[102]

Matoba
T.
Shimokawa
H.
Nakashima
M.
Hirakawa
Y.
Mukai
Y.
Hirano
K.
et al. 

Hydrogen peroxide is an endothelium-derived hyperpolarizing factor in mice
J Clin Invest
2000
106
1521
1530
[103]

Yada
T.
Shimokawa
H.
Hiramatsu
O.
Kajita
T.
Shigeto
F.
Goto
M.
et al. 

Hydrogen peroxide, an endogenous endothelium-derived hyperpolarizing factor, plays an important role in coronary autoregulation in vivo
Circulation
2003
107
1040
1045
[104]

Iida
Y.
Katusic
Z.S.

Mechanisms of cerebral arterial relaxations to hydrogen peroxide
Stroke
2000
31
2224
2230
[105]

Miura
H.
Bosnjak
J.J.
Ning
G.
Saito
T.
Miura
M.
Gutterman
D.D.

Role for hydrogen peroxide in flow-induced dilation of human coronary arterioles
Circ Res
2003
92
e31
e40
[106]

Thengchaisri
N.
Kuo
L.

Hydrogen peroxide induces endothelium-dependent and -independent coronary arteriolar dilation: role of cyclooxygenase and potassium channels
Am J Physiol Heart Circ Physiol
2003
285
H2255
H2263
[107]

Cosentino
F.
Barker
J.E.
Brand
M.P.
Heales
S.J.
Werner
E.R.
Tippins
J.R.
et al. 

Reactive oxygen species mediate endothelium-dependent relaxations in tetrahydrobiopterin-deficient mice
Arterioscler Thromb Vasc Biol
2001
21
496
502
[108]

van Hinsbergh
V.W.

NO or H(2)O(2) for endothelium-dependent vasorelaxation: tetrahydrobiopterin makes the difference
Arterioscler Thromb Vasc Biol
2001
21
719
721
[109]

Wolin
M.S.
Burke
T.M.

Hydrogen peroxide elicits activation of bovine pulmonary arterial soluble guanylate cyclase by a mechanism associated with its metabolism by catalase
Biochem Biophys Res Commun
1987
143
20
25
[110]

Iesaki
T.
Gupte
S.A.
Kaminski
P.M.
Wolin
M.S.

Inhibition of guanylate cyclase stimulation by NO and bovine arterial relaxation to peroxynitrite and H2O2
Am J Physiol
1999
277
H978
H985
[111]

Ellis
A.
Pannirselvam
M.
Anderson
T.J.
Triggle
C.R.

Catalase has negligible inhibitory effects on endothelium-dependent relaxations in mouse isolated aorta and small mesenteric artery
Br J Pharmacol
2003
140
1193
1200
[112]

Fujimoto
S.
Mori
M.
Tsushima
H.

Mechanisms underlying the hydrogen peroxide-induced, endothelium-independent relaxation of the norepinephrine-contraction in guinea-pig aorta
Eur J Pharmacol
2003
459
65
73
[113]

Drummond
G.R.
Cai
H.
Davis
M.E.
Ramasamy
S.
Harrison
D.G.

Transcriptional and posttranscriptional regulation of endothelial nitric oxide synthase expression by hydrogen peroxide
Circ Res
2000
86
347
354
[114]

Cai
H.
McNally
J.S.
Weber
M.
Harrison
D.G.

Oscillatory shear stress upregulation of endothelial nitric oxide synthase requires intracellular hydrogen peroxide and CaMKII
J Mol Cell Cardiol
2004
37
121
125
[115]

Kanazawa
K.
Kawashima
S.
Mikami
S.
Miwa
Y.
Hirata
K.
Suematsu
M.
et al. 

Endothelial constitutive nitric oxide synthase protein and mRNA increased in rabbit atherosclerotic aorta despite impaired endothelium-dependent vascular relaxation
Am J Pathol
1996
148
1949
1956
[116]

Hink
U.
Li
H.
Mollnau
H.
Oelze
M.
Matheis
E.
Hartmann
M.
et al. 

Mechanisms underlying endothelial dysfunction in diabetes mellitus
Circ Res
2001
88
E14
E22
[117]

Bradley
J.R.
Johnson
D.R.
Pober
J.S.

Endothelial activation by hydrogen peroxide. Selective increases of intercellular adhesion molecule-1 and major histocompatibility complex class I
Am J Pathol
1993
142
1598
1609
[118]

Ziegelstein
R.C.
He
C.
Hu
Q.

Hypoxia/reoxygenation stimulates Ca2+-dependent ICAM-1 mRNA expression in human aortic endothelial cells
Biochem Biophys Res Commun
2004
322
68
73
[119]

Lakshminarayanan
V.
Lewallen
M.
Frangogiannis
N.G.
Evans
A.J.
Wedin
K.E.
Michael
L.H.
et al. 

Reactive oxygen intermediates induce monocyte chemotactic protein-1 in vascular endothelium after brief ischemia
Am J Pathol
2001
159
1301
1311
[120]

Chen
X.L.
Zhang
Q.
Zhao
R.
Ding
X.
Tummala
P.E.
Medford
R.M.

Rac1 and superoxide are required for the expression of cell adhesion molecules induced by tumor necrosis factor-alpha in endothelial cells
J Pharmacol Exp Ther
2003
305
573
580
[121]

Chen
X.L.
Zhang
Q.
Zhao
R.
Medford
R.M.

Superoxide, H2O2, and iron are required for TNF-alpha-induced MCP-1 gene expression in endothelial cells: role of Rac1 and NADPH oxidase
Am J Physiol Heart Circ Physiol
2004
286
H1001
H1007
[122]

Takaishi
H.
Taniguchi
T.
Takahashi
A.
Ishikawa
Y.
Yokoyama
M.

High glucose accelerates MCP-1 production via p38 MAPK in vascular endothelial cells
Biochem Biophys Res Commun
2003
305
122
128
[123]

True
A.L.
Rahman
A.
Malik
A.B.

Activation of NF-kappaB induced by H(2)O(2) and TNF-alpha and its effects on ICAM-1 expression in endothelial cells
Am J Physiol Lung Cell Mol Physiol
2000
279
L302
L311
[124]

Pueyo
M.E.
Gonzalez
W.
Nicoletti
A.
Savoie
F.
Arnal
J.F.
Michel
J.B.

Angiotensin II stimulates endothelial vascular cell adhesion molecule-1 via nuclear factor-kappaB activation induced by intracellular oxidative stress
Arterioscler Thromb Vasc Biol
2000
20
645
651
[125]

Lewis
M.S.
Whatley
R.E.
Cain
P.
McIntyre
T.M.
Prescott
S.M.
Zimmerman
G.A.

Hydrogen peroxide stimulates the synthesis of platelet-activating factor by endothelium and induces endothelial cell-dependent neutrophil adhesion
J Clin Invest
1988
82
2045
2055
[126]

Okayama
N.
Coe
L.
Oshima
T.
Itoh
M.
Alexander
J.S.

Intracellular mechanisms of hydrogen peroxide-mediated neutrophil adherence to cultured human endothelial cells
Microvasc Res
1999
57
63
74
[127]

Fraticelli
A.
Serrano
C.V.
Jr.
Bochner
B.S.
Capogrossi
M.C.
Zweier
J.L.

Hydrogen peroxide and superoxide modulate leukocyte adhesion molecule expression and leukocyte endothelial adhesion
Biochim Biophys Acta
1996
1310
251
259
[128]

Shappell
S.B.
Toman
C.
Anderson
D.C.
Taylor
A.A.
Entman
M.L.
Smith
C.W.

Mac-1 (CD11b/CD18) mediates adherence-dependent hydrogen peroxide production by human and canine neutrophils
J Immunol
1990
144
2702
2711
[129]

Hong
Y.H.
Peng
H.B.
La Fata
V.
Liao
J.K.

Hydrogen peroxide-mediated transcriptional induction of macrophage colony-stimulating factor by TGF-beta1
J Immunol
1997
159
2418
2423
[130]

Okada
H.
Woodcock-Mitchell
J.
Mitchell
J.
Sakamoto
T.
Marutsuka
K.
Sobel
B.E.
et al. 

Induction of plasminogen activator inhibitor type 1 and type 1 collagen expression in rat cardiac microvascular endothelial cells by interleukin-1 and its dependence on oxygen-centered free radicals
Circulation
1998
97
2175
2182
[131]

Lang
I.M.
Moser
K.M.
Schleef
R.R.

Elevated expression of urokinase-like plasminogen activator and plasminogen activator inhibitor type 1 during the vascular remodeling associated with pulmonary thromboembolism
Arterioscler Thromb Vasc Biol
1998
18
808
815
[132]

Kaikita
K.
Schoenhard
J.A.
Painter
C.A.
Ripley
R.T.
Brown
N.J.
Fogo
A.B.
et al. 

Potential roles of plasminogen activator system in coronary vascular remodeling induced by long-term nitric oxide synthase inhibition
J Mol Cell Cardiol
2002
34
617
627
[133]

Weisberg
A.D.
Albornoz
F.
Griffin
J.P.
Crandall
D.L.
Elokdah
H.
Fogo
A.B.
et al. 

Pharmacological inhibition and genetic deficiency of plasminogen activator inhibitor-1 attenuates angiotensin II/salt-induced aortic remodeling
Arterioscler Thromb Vasc Biol
2004
25
365
371
[134]

Galis
Z.S.
Khatri
J.J.

Matrix metalloproteinases in vascular remodeling and atherogenesis: the good, the bad, and the ugly
Circ Res
2002
90
251
262
[135]

Yang
H.
Roberts
L.J.
Shi
M.J.
Zhou
L.C.
Ballard
B.R.
Richardson
A.
et al. 

Retardation of atherosclerosis by overexpression of catalase or both Cu/Zn-superoxide dismutase and catalase in mice lacking apolipoprotein E
Circ Res
2004
95
1075
1081
[136]

Tribble
D.L.
Gong
E.L.
Leeuwenburgh
C.
Heinecke
J.W.
Carlson
E.L.
Verstuyft
J.G.
et al. 

Fatty streak formation in fat-fed mice expressing human copper-zinc superoxide dismutase
Arterioscler Thromb Vasc Biol
1997
17
1734
1740
[137]

Jung
O.
Marklund
S.L.
Geiger
H.
Pedrazzini
T.
Busse
R.
Brandes
R.P.

Extracellular superoxide dismutase is a major determinant of nitric oxide bioavailability: in vivo and ex vivo evidence from ecSOD-deficient mice
Circ Res
2003
93
622
629
[138]

Fennell
J.P.
Brosnan
M.J.
Frater
A.J.
Hamilton
C.A.
Alexander
M.Y.
Nicklin
S.A.
et al. 

Adenovirus-mediated overexpression of extracellular superoxide dismutase improves endothelial dysfunction in a rat model of hypertension
Gene Ther
2002
9
110
117
[139]

Sentman
M.L.
Brannstrom
T.
Westerlund
S.
Laukkanen
M.O.
Yla-Herttuala
S.
Basu
S.
et al. 

Extracellular superoxide dismutase deficiency and atherosclerosis in mice
Arterioscler Thromb Vasc Biol
2001
21
1477
1482

Author notes

Time for primary review 27 days