Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Jan 1.
Published in final edited form as: Front Biosci (Schol Ed). 2012 Jan 1;4:1044–1064. doi: 10.2741/s317

Differential roles of NADPH oxidases in vascular physiology and pathophysiology

Angelica M Amanso 1, Kathy K Griendling 1
PMCID: PMC3358302  NIHMSID: NIHMS377395  PMID: 22202108

Abstract

Reactive oxygen species (ROS) are produced by all vascular cells and regulate the major physiological functions of the vasculature. Production and removal of ROS are tightly controlled and occur in discrete subcellular locations, allowing for specific, compartmentalized signaling. Among the many sources of ROS in the vessel wall, NADPH oxidases are implicated in physiological functions such as control of vasomotor tone, regulation of extracellular matrix and phenotypic modulation of vascular smooth muscle cells. They are involved in the response to injury, whether as an oxygen sensor during hypoxia, as a regulator of protein processing, as an angiogenic stimulus, or as a mechanism of wound healing. These enzymes have also been linked to processes leading to disease development, including migration, proliferation, hypertrophy, apoptosis and autophagy. As a result, NADPH oxidases participate in atherogenesis, systemic and pulmonary hypertension and diabetic vascular disease. The role of ROS in each of these processes and diseases is complex, and a more full understanding of the sources, targets, cell-specific responses and counterbalancing mechanisms is critical for the rational development of future therapeutics.

Keywords: Reactive oxygen species, NADPH oxidases, Nox, Vascular disease, Review

2. INTRODUCTION

Virtually all resident cells in the vessel wall produce and respond to reactive oxygen species (ROS) (14). Vascular smooth muscle cells (VSMCs) in response to stimuli change phenotype from quiescent/contractile to synthetic, being able to proliferate, migrate and synthesize extracellular matrix. Endothelial cells respond to vasoactive agents by releasing nitric oxide (NO) to regulate vascular tone, by maintaining the pro/antithrombotic balance, and by promoting or inhibiting permeability and inflammation. Adventitial fibroblasts can influence vessel contraction and medial hypertrophy. ROS have been recognized as key regulators of many of these responses. More recently, it has been recognized that redox processes occur in a compartmentalized manner, with generation of different ROS in distinct organelles by several cellular enzyme systems (57). Vascular cells have diverse sources of ROS, namely from oxidative ATP generation in the mitochondria, xanthine oxidase during the metabolic pathway for acid uric formation, oxidoreductases such as cytochrome P450, glutathione peroxidase and myeloperoxidase, multiple thiol-disulfide exchanges during folding of nascent proteins in the endoplasmic reticulum, uncoupled nitric oxide synthases and Nox family NADPH oxidases (3, 5, 812). Maintaining the proper local balance of ROS production and removal is essential for normal physiology, with excess ROS potentially contributing to disease development (13). This review will focus on the role of ROS in both of these processes, with particular emphasis on NADPH oxidase-derived signaling.

3. NADPH OXIDASES

The Nox protein family is a major source of ROS in vascular cells (8, 14). The first member of this family to be described was gp91phox or Nox2, which generates large amounts of ROS in phagocytic cells, acting as an important component of innate immunity (15). In 1999, mox1 or NOH-1 (now called Nox1), a homolog of gp91phox, was discovered and was identified as an enzymatic source of ROS in non-phagocyte cells (16, 17). Since then seven members of the Nox family have been identified (Nox1-Nox5, Duox1 and Duox2) (2, 18). These isoforms differ from one another in their patterns of expression, subcellular localization, type of ROS produced and mode of activation (Table 1).

Table 1.

Nox enzymes in the vasculature

nox member localization ROS type regulators expression activation cardiovascular physiology cardiovascular pathophysiology
Nox1 transmenbrane
caveolae
endosomes
superoxide anion p22phox
p47phox, Noxa1, Rac or Noxo1, Noxa1, Rac
smooth muscle cells
epithelial cells, uterus
osteoclasts, neurons
stimulus induced
Ang II, PDGF, FGF
proliferation, migration
extracellular matrix production
host defense
vascular hypertrophy
hypertension, vascular injury
diabetes
Nox2 phagosomes
caveolae
endosomes
superoxide anion p22phox, p47phox, p67phox, p40phox
rac
endothelial cells
cardiomyocytes, hepatocytes
fibroblastes, phagocytes
neurons, skeletal muscle myocytes
agonist-induced (Ang II, VEGF, TNFα, endothelin1)
mechanical forces
angiogenesis
apoptosis, oxygen sensing
inflammation
hypoxia, diabetes
hypercholesterolemia
pulmonary hypertension
Nox3 plasma membrane superoxide anion p22phox, Noxo1
possibly Noxa1, Rac
inner ear
lung endothelial cells
hepatocytes
agonist-induced (TNFα)
constitutively active
unknown insulin resistance
Nox4 focal adhesion
nucleus
endoplasmic reticulum
hydrogen peroxide p22phox
Poldip2
mesangial cells, smooth muscle
endothelial cells, fibroblasts
keratinocytes, osteoclasts
neurons, hepatocytes
cardiomyocytes
constitutively active
stimulus induced (TGFβ, IL-1, Vit.D, thrombin, ER stressors)
senescence, apoptosis
endoplasmic reticulum stress
survival, differentiation
oxygen sensing
cytoskeletal regulation
hypertrophy
pulmonary hypertension/fibrosis
metabolic syndrome
Nox5 internal membranes
plasma membrane
superoxide calcium endothelial cells
smooth muscle cells
testes, spleen
calcium activated proliferation
inflammation
unknown

The Nox proteins generate ROS by transfer of a single electron from NADPH to molecular oxygen (18). The proteins have an oxidoreductase structure with a transmembrane region that contains two heme groups and a cytosolic portion with NADPH and FAD binding sites (8, 1820). Nox1-4 require p22phox for their activity while Nox5 does not (2022). p22phox is a regulatory membrane protein that functions as a docking site for regulatory proteins (p40phox, p47phox and p67phox for Nox2; Noxa1 and Noxo1 for Nox1 and 3; polymerase delta interacting protein 2 (Poldip2) for Nox4) (2, 10, 2327). The small GTPase Rac also regulates Nox1-3 activity (8, 23, 2830). Nox5 also differs from the other Nox enzymes in that it does not require any cytosolic subunits (31). Structure predicts that all isoforms release superoxide (O2·−); however, while this is true for Nox1, Nox2, Nox3 and Nox5, Nox4 and the Duox enzymes release mostly hydrogen peroxide (H2O2) (8, 32). In vascular cells, Nox1, Nox2, Nox4 and Nox5 have been linked to several pathophysiological conditions such as hypertension, atherosclerosis, restenosis after injury and diabetic vasculopathy (3236).

Nox1, found in plasma membrane, caveolae and endosomes in VSMCs, is induced by growth factors and vasoactive agents and activated by Src-dependent, protein kinase C (PKC)-mediated p47phox phosphorylation (18, 37, 38). Nox2 is present in plasma membrane and is activated similarly to vascular Nox1 (14, 39). Nox4 is thought to be constitutively activated and its activity is correlated with its level of expression (40). However, recent studies show that Nox4 is regulated by Poldip2 and such regulation increases its activity, although the mechanism remains to be determined (24). This novel partner colocalizes with p22phox and Nox4 at focal adhesions and in the nucleus of VSMCs (24). Nox4 is also found in the perinuclear space, endoplasmic reticulum and plasma membrane (37). Another potential regulatory mechanism for Nox4 (and Nox1) involves a chaperone of the endoplasmic reticulum, protein disulfide isomerase (PDI) (41). Downregulation of PDI decreases the Nox activity induced by angiotensin II (Ang II), which has previously been attributed to Nox1 (42). For Nox4, however, the relationship with PDI may be due to an interaction between oxidative and endoplasmic reticulum stress (4345). Nox5 is regulated by elevation of calcium levels, and is found in the cytoskeletal fraction, endoplasmic reticulum and plasma membrane. Like Nox5, Duox enzymes are activated directly by calcium, although they do not appear to be expressed in the vasculature (8, 10, 14, 18, 20, 22, 46).

4. REGULATION OF VASCULAR TONE

Nox-derived ROS have physiological and pathophysiological roles in the cardiovascular system. As noted, Nox enzymes promote cell growth, migration, proliferation and differentiation, and these effects have been related with pathophysiological conditions. However, Nox enzymes are not important only in cell injury conditions, but also during development, reparative processes and in the physiology of the vascular tone and oxygen sensing (8, 47).

The endothelial regulation of vascular tone is directly correlated with local production and activity of NO. This regulation is sensitive to any alterations in local stimuli such as shear stress or the presence of local vasodilators. In response to these stimuli, increases and redistribution of blood flow occur in specific vascular networks (4850). The presence of O2·− in this environment is an important foil to NO action, because O2·− can rapidly interact with NO to form peroxynitrite (51, 52). Consequently, the bioavailability of NO is reduced and other reactive species are generated. Peroxynitrite can 1) uncouple endothelial nitric oxide synthase (eNOS) by oxidizing tetrahydrobiopterin (BH4) thus leading to reduced NO and increased O2·−; 2) inactivate prostacyclin synthase (an important source of endothelium-derived relaxing factor); and 3) inhibit superoxide dismutases, inducing oxidative stress (53, 54).

Conversely, it has been shown that H2O2 can induce and activate eNOS (55, 56). H2O2 can also exert vasodilatory effects in vascular cells by hyperpolarization mechanisms (5559). Recently, endothelial H2O2 has been proposed to be the endothelium-dependent hyperpolarization factor (EDHF) stimulated by shear stress to induce relaxation of the vessel (60, 61). Moreover, vascular overexpression of catalase counteracts the activity of some vasoconstrictor agents, suggesting that endogenous H2O2 may be an important regulator of blood pressure (62). On the other hand, other studies report that catalase has minimal effects on endothelium dependent relaxation in aorta and small arteries (63). Adventitia-derived H2O2 has been proposed to impair vascular relaxation via p38 mitogen-activated protein kinase (MAPK) activation and SH2 domain–containing protein tyrosine phosphatase–2 (SHP2) inhibition. Other findings implicate H2O2-dependent extracellular signal regulated kinase (ERK) 1/2 activation in inhibiting relaxation (58, 64, 65).

It is thus clear that ROS critically regulate vascular tone, but that their effects may be dependent upon the identity of ROS, the nature of the stimulus and the vascular bed upon which it acts. For example, H2O2 relaxes coronary and pulmonary arteries but may relax or constrict mesenteric arteries, while O2·− inactivates NO to counteract relaxation and may directly constrict vessels (66). In human coronary arterioles, bradykinin-induced dilation is reduced by the Nox2 inhibitor gp91ds-tat, suggesting that Nox2 mediates this response, but the relevant ROS is H2O2, which reduces the bioavailability of epoxyeicosatrienoic acids (67). Moreover, in this arterial bed, flow induces endothelial production of H2O2, which activates large conductance Ca2+-activated K+ channels on the smooth muscle cells to induce relaxation (59). The multiple roles of ROS complicate the potential usefulness of Nox inhibitors in blood pressure regulation, since reduction of O2·− would favor improved relaxation and thus a lowering of blood pressure, but reduction of H2O2 might have the unwanted side effect of impairing coronary vasodilation. Because Nox1/2 generate O2·` while Nox4 releases mostly H2O2, identifying which Nox is responsible for which response may allow us to specifically target the offending Nox in specific diseases.

5. VASCULAR REMODELING AND EXTRACELLULAR MATRIX REORGANIZATION

Vascular remodeling refers to a structural modification that results in changes in wall thickness and lumen diameter. Factors that may induce this response include passive adaptation to chronic hemodynamic stimuli, growth factors such as Ang II, and ROS. There are two kinds of remodeling: eutrophic or hypertrophic. In eutrophic remodeling the luminal diameter is reduced and the media becomes thicker with no overall change in cross-sectional area. However, in the hypertrophic scenario wall cross-sectional area is increased, cell size is increased, and extracellular matrix (ECM) accumulation of proteins such as collagen and fibronectin is enhanced (6871). This latter type of remodeling is present in hypertension (72).

Findings in the literature suggest an important role of Nox-derived ROS in hypertrophic remodeling. New Zealand obese mice with metabolic syndrome have increased vascular O2·− and peroxynitrite production, Nox activity and adhesion molecule expression, leading to hypertrophic vascular remodeling in mesenteric resistance arteries characterized by increased media:lumen ratio and media cross-sectional area (73, 74). Systemic administration of gp91phox-dstat, an inhibitor of Nox2, decreases Ang II-induced vascular hypertrophy (75). Nox1 has also been implicated in the oxidative, pressor, and hypertrophic responses to Ang II (76, 77).

Vascular remodeling also requires degradation and reorganization of the ECM proteins. The most important enzymes involved in these spatial dynamics are matrix metalloproteinases, a group of zinc-dependent endopeptidases that degrade ECM proteins such as fibronectin and collagen. High expression/activation of MMPs is a critical component of the vascular remodeling process (78, 79). Several studies indicate a role of Nox-derived ROS in ECM turnover. ROS promote activation of MMP-2 and MMP-9 in VSMCs (80, 81). In addition, Ang II induces Nox-dependent expression of MMP-1, MMP-3 and MMP-9 (82). Mechanical stretch stimulates expression and release of MMP-2 by Nox-derived ROS (83). Moreover, the chloride-proton antiporter chloride channel-3 (ClC-3), which is required for endosome-dependent signaling by Nox1 in VSMCs, contributes to neointimal hyperplasia after vascular injury by TNFalpha-mediated activation of MAPKs and MMP-9 (8486). Similar effects of ROS on MMP-induced remodeling occur in other tissues as well, including TGFbeta-induced degradation of ECM and increased activity of MMP-9 during metastasis (71), TGFbeta-induced, Nox4-dependent fibrogenesis in lung injury (87), and TNFalpha-mediated, Nox-derived ROS-dependent expression of MMP-9, but not MMP-2, in cardiomyocytes (88).

Lipopolysaccharide (LPS) induces ROS generation and increases the expression of Nox1 and Nox2, as well as the expression of several MMPs. Recent work showed that the regulation of MMPs and migration in macrophages stimulated with LPS is dependent on Nox2-, but not Nox1-derived ROS. While Nox1 small interfering RNA (siRNA) did not inhibit LPS-mediated expression of MMPs, Nox2 siRNA inhibited the expression of MMP-9, 10 and 12 (89).

6. PHENOTYPIC MODULATION (PROLIFERATION AND DIFFERENTIATION)

ROS including O2·− and H2O2 have important roles in the regulation of both proliferation and differentiation (39, 90). These mechanisms are crucial for not only the control of vasomotor tone, but also tissue repair and vascular remodeling. Low concentrations of O2·− and H2O2 induce proliferation, moderate concentrations induce either cell stasis or hypertrophy, and very high concentrations can induce apoptosis (39, 9195). Thus, as with other cellular functions, the amount of ROS produced, their source and the subcellular compartment in which they are generated greatly influence the final response.

Differentiated VSMCs express contractile proteins such as smooth muscle α-actin, calponin and smooth muscle myosin heavy chain, and have low proliferative capability and synthetic activity. In cardiovascular pathologies, VSMCs switch to the synthetic phenotype, proliferate and migrate, predisposing to vascular hypertrophy and lesion formation with elevated production of extracellular matrix (90). The mechanisms behind this phenotype switching are still very obscure; however, it is clear that physical and chemical changes in the environment modulate this process, namely humoral factors (platelet-derived growth factor (PDGF), TGFbeta, Ang II, endothelin-1 (ET-1)), neuronal signals, mechanical injury and changes in mechanical forces affecting cell-cell interaction and extracellular matrix (90). Nox1 and Nox4 appear to play very different roles during phenotypic modulation.

Nox4 is crucial for the maintenance of differentiation marker gene expression in VSMCs, stem cells, fibroblasts and pulmonary artery smooth muscle cells (14, 9699). Low serum conditions or TGFbeta have been used as an external signal to induce differentiation in VSMC. A recent study using TGFbeta suggests that Nox4 regulates smooth muscle alpha-actin gene expression via activation of p38MAPK and the serum-response factor (SRF)/myocardin-related transcription factor (MRTF) pathway in response to TGFbeta. Importantly, Nox4 has also been implicated in the differentiation of adipocytes, osteoblasts, chondrocytes and myofibroblasts. In adipocytes, Nox4 acts as a switch from insulin-induced proliferation to differentiation by controlling MAPK phosphatase-1 (MKP-1) expression, which controls ERK1/2 signaling (100). Similar to VSMCs, Nox4 mediates TGFbeta-stimulated fibroblast differentiation into a profibrotic, myofibroblast phenotype as well as the accompanying matrix production (101).

In contrast to the role of Nox4 in differentiation, Nox-mediated cell proliferation has been described in a several cell types present in blood vessels, kidney, liver, heart, lung epithelial cells and wide range of cancer cells (39, 102105). In the blood vessel wall, ROS-mediated proliferation of VSMCs, adventitial fibroblasts and endothelial cells has been shown to be dependent on the Nox family proteins (105107). Importantly, Nox isoforms seem to have different roles depending on their subcellular compartmentalization (8, 14).

Nox1 is fundamentally important in cell proliferation in VSMCs. Nox1 is located in caveolae and downregulation of caveolin decreased ROS production as well cell proliferation in VSMCs treated with Ang II (37, 84, 108, 109). In addition, it has been implicated in the growth response to PDGF, thrombin and ET-1. Overexpression of Nox1 potentiates the effect of Ang II on vascular hypertrophy while downregulation inhibits it (103, 110112). The importance of Nox1 in proliferation has also been established in studies of neointima formation using genetically modified animal models (108), where injury-induced neointima formation was reduced in Nox1 deficient animals and proliferation and migration were reduced in VSMCs cultured from these animals. Conversely, overexpression of the Nox1 activator NoxA1 in mouse carotid arteries increased O2·− production in VSMCs and enhanced neointimal formation and atherosclerotic lesion formation (113). The mechanism appears to involve thrombin-induced activation of redox-sensitive mitogenic protein kinases that lead to proliferation.

The molecular mechanisms controlling Nox1-mediated proliferation are only partially understood, but include activation of pro-growth signaling molecules such as kinases (p38MAPK and Akt), Ras and redox-sensitive transcriptional factors and inhibition of protein tyrosine phosphatases (112, 114). Only a few direct targets of ROS have been identified, and these are usually proteins with low pKa cysteine residues that can be oxidized by H2O2 (115, 116). Thus, antioxidant enzymes such as compartmentalized peroxiredoxins or thioredoxin are important regulators of the overall response (116).

Other recent findings underline the importance of cross-talk between redox pathways in Nox1-dependent proliferation. Manipulating the extracellular oxidation-reduction state in aortic segments and VSMCs by altering the cysteine concentration (117) leads to an upregulation of Nox1 and an increase in proliferation. The mechanism appears to involve epidermal growth factor receptor phosphorylation by cleaved extracellular ligands following by ERK 1/2 activation.

In endothelial cells, Nox2 and Nox4 are the homologues that participate in cell proliferation. Their roles are not additive, but may be partially redundant (118120). Both EaHy926 endothelial cells (a cell line) and human microvascular endothelial cells utilize Nox2 and Nox4 equally for ROS production and proliferation. Depletion of Nox1 in these cells does not alter ROS levels under basal conditions (120). Another study showed that overexpression of Nox4 enhanced receptor tyrosine kinase phosphorylation and the activation of ERK1/2, leading to enhanced proliferation and migration of endothelial cells (121).

Nox2 has also been described as an important regulator of fibroblast proliferation since its downregulation results in reduction of serum-induced proliferation (122, 123). Others findings suggest that ROS-derived from Nox family proteins in fibroblasts act as a paracrine stimulus on intimal cells and VSMCs to activate MAPKs and induce hypertrophy (124). Recently, it has been shown that Nox2 can regulate intracellular growth signaling pathways in adult hippocampal stem/progenitor cells, suggesting that, at least in the brain, ROS are important for maintaining specific cell populations and are not just related with deleterious conditions (125).

7. MIGRATION

Migration is a natural cellular process that is an integral part of vascular development as well as the healing response after vascular injury (126). ROS were first implicated in vascular migration more than 15 years ago when Sundarasen et al (104) showed that PDGF-induced VSMC migration was inhibited by catalase. Direct involvement of Nox-derived ROS in migration has also been shown for Ang II, ox-LDL, hypoxia, PDGF and vascular endothelial growth factor (VEGF) stimulation of either VSMCs or endothelial cells (108, 127131). The signaling molecules so far identified as being regulated by ROS during migration are Src, RhoA, protein kinase C, phosphoinositide-3 (PI3) kinase, MAPKs including ERK1/2 and p38MAPK, NFkappaB, cofilin, slingshot-1 phosphatase (SSH1L) and Akt (126).

Several groups have described an interrelationship between PDGF and Nox1 or between basic fibroblast growth factor (bFGF) and Nox1 during migration (108, 127, 128, 132, 133). Other work showed that heme oxygenase (HO-1) and carbon monoxide (CO) inhibit PDGF stimulated VSMC migration via inhibition of Nox1 activity (134). The mechanisms mediating the role of Nox1 in migration include the modulation of cytoskeletal rearrangement as well as regulation of matrix metalloproteinases. In terms of cytoskeletal dynamics, PDGF regulates SSH1L phosphatase, an enzyme responsible for activation of cofilin and actin depolymerization, through Nox1-derived ROS (135). Similarly, bFGF activates c-Jun N-terminal kinase (JNK) via Nox1-dependent ROS production (136). Nox1 siRNA blocks the up-regulation of MMP-9 in K-Ras-transformed normal rat kidney cells. Downregulation of Nox1 inhibits the activation of inhibitor of NFkappaB kinase-alpha (IKKalpha) and the degradation of IkappaBalpha, thus suppressing the NFkappaB activity in the promoter of MMP-9 (137). In other models of migration, Nox1 knockdown inhibits the directional persistence of migration via Rho-dependent modulation of integrin expression (138).

Nox4 also appears to have an important function in migration (139). Knockdown of Nox4 inhibits insulin-like growth factor-1 (IGF-1)-induced migration at least in part due to suppression of MMP2 and MMP9 induction (139). PDGF-induced migration is inhibited by siRNA against Nox4 or its regulator Poldip2 by a mechanism involving focal adhesion turnover (24). In a co-culture system of stromal mammary cells and epithelial cells, Nox4-dependent ROS production by TGFbeta stimulation of stromal cells mediates epithelial cell migration (140). Conversely, Nox4 overexpression has been shown to inhibit Ang II-induced myofibroblast migration via an H2O2 dependent pathway (141).

Normally, NO inhibits VSMC migration. However, in smooth muscle cells exposed to high glucose, NO loses its ability to inhibit migration due to glucose-induced oxidation of the sarco/endoplasmic reticulum ATPase (SERCA) (142). Both Nox1 and Nox4 are upregulated in these conditions, and knockdown of either oxidase restores the inhibition of migration by NO. Blocking activation of Rac1 or p47phox has the same effect. Thus, it appears that the Nox family is the source of the oxidants responsible for the failure of NO to inhibit SMC migration in high glucose conditions (142). Similar results have been described in cultured aortic SMC from obese Zucker rats, a model of obesity and insulin resistance. Upregulation of Nox4 by TGFbeta via Smad2 promotes oxidation of SERCA and consequently impairs NO-mediated inhibition of serum-induced VSMC migration (143).

8. APOPTOSIS AND AUTOPHAGY

Cell survival/growth and cell death are cellular responses integral to all physiological processes. However, in many pathological conditions cell death induces tissue dysfunction leading to disease. Cells from renewable tissues have two mechanisms to avoid uncontrolled growth, namely cellular senescence, which consists of permanent cell cycle arrest, or initiation of cell death programs (apoptosis or autophagy). Apoptosis is defined as programmed cell death while autophagy represents a lysosomal degradation pathway of cytoplasmic components such as protein aggregates and damaged organelles. The primary function of autophagy is to maintain cellular homeostasis. Cells already terminally differentiated, such as neurons and cardiomyocytes, cannot become senescent (144, 145), but are subject to both apoptosis and autophagy.

ROS have been linked to cell death extensively in the literature. It is believed that temporally and spatially produced ROS regulate specific redox signaling pathways, while prolonged ROS generation in a diffuse manner can induce oxidative damage and consequently cell death (146, 147). Because Nox enzymes have specific subcellular localizations, they are likely to regulate the more specific death pathways. It is known, for example, that ROS regulate pro-apoptotic signaling (p38MAPK; JNK; apoptosis signal-regulating kinase (ASK1); Ca2+/calmodulin-dependent protein kinase II (CaMKII); C/EBP-homologous protein (CHOP)) as well as anti-apoptotic pathways (Akt, ERK1/2, and heat shock proteins (HSPs)) (146).

Apoptosis is induced in vascular endothelial cells by homocysteine, TNFalpha and remnant lipoprotein particles (148150). In addition, UV radiation promotes activation of Nox with consequent increased apoptosis in keratinocytes (151). Studies in the literature have implicated Nox2 in cardiomyocyte apoptosis induced by Ang II. In the presence of apocynin or in p47phox knockout mice apoptosis is abolished, suggesting the involvement of Nox2 (Nox1 and 3 are not detected in these cells) (152). Nox4 also appears to induce apoptosis in cultured cardiomyocytes as well mouse heart tissue. Nox4-induced apoptosis is accompanied by cytochrome-c release and is reversed by Bcl-xl and manganese superoxide dismutase (MnSOD), suggesting involvement of the mitochondrial pathway. In addition, cardiac-specific Nox4 knockout mice have attenuated cardiomyocyte apoptosis (153, 154).

It is known that inactivation of the tumor suppressor p53 is strongly associated with pathological conditions in cardiovascular diseases. Overexpression of p53 promotes an increase in ROS in VSMCs with inhibition of proliferation and induction of apoptosis (102, 155157), indicating that cell-cycle progression regulated by p53 in VSMCs is negatively regulated by ROS. Moreover, it has been described that VEGF-B is a potent negative regulator of apoptosis, potentially by inhibiting the expression of proapoptotic BH3-only proteins (members of the proapoptotic branch of the Bcl-2 proteins) in mouse primary VSMCs (158).

Another mechanism leading to apoptosis is endoplasmic reticulum (ER) stress. This is a type of cellular dysfunction in which ER function, especially protein synthesis, is compromised, generating a specific signaling response called the Unfolded Protein Response (UPR) in order to restore homeostasis (159, 160). Transient ER stress is part of normal cellular physiology; however, chronic ER stress contributes to pathological conditions such as atherosclerosis, cardiac hypertrophy and the response to ischemia. If UPR fails to restore ER function, ER-initiated apoptotic signaling is induced (161163).

In recent years, an important convergence between oxidative stress and ER stress has been recognized (164). Initially, it was thought that oxidative stress resulted from ER stress upon redox folding of nascent proteins where ROS generation occurs during thiol-disulfide exchanges involving ER oxidoreductases, including the flavooxidase Ero1 and PDI (165). However, a range of studies has correlated ER stress with Nox family proteins (43, 166173). In human VSMCs, Nox4 mediates apoptosis induced by 7-ketocholesterol. The mechanisms involved are related with the apoptotic branch of ER stress signaling, namely the CHOP signaling pathway and activation of inositol-requiring enzyme-1 (IRE1), a proximal sensor for the status of ER luminal protein folding during the UPR (172). Moreover, a recent study showed that proteasome inhibition during UPR can directly affect Nox4 activity and impairs VSMC viability during ER stress induced by tunicamycin (166). Another important role of Nox4 in ER stress is related to its functional interaction with the phosphatase PTP1B in the ER. PTP1B is involved in the IRE-1 branch of UPR signaling during ER stress (174). Oxidation of PTP1B by ROS-derived from Nox4 reduces its activity (175). Finally, a recent study demonstrated that ischemia/reperfusion injury induces Nox4 expression in renal tubule cells. The authors showed an interaction between Toll-like receptor 4, a 28 kDa form of Nox4 and the ER resident protein gp96, a homologue of cytosolic Hsp90 that is involved in ROS production and TNF receptor-associated factor 2 (TRAF2)/ASK1/JNK-mediated apoptosis in posthypoxic renal tubule cells. The authors suggested that Nox4 acts as a sensor in ER stress–induced proapoptotic signaling in this system (176).

Recent work showed that Nox2 also has an important role in oxidative and ER stress-induced apoptosis. Li et al (168) found that a branch of UPR signaling involving CHOP, ERO1alpha and CAMKII induces Nox2 and oxidative stress, which are necessary for ER stress induced apoptosis. In in vivo experiments, Nox2 deficiency protects against ER stress-induced renal cell apoptosis and CHOP activation as well renal dysfunction.

ER stress and autophagy are completely inter-related processes (163). It has been described that ER stress induces autophagy to protect against cell death (163). In addition, in the last few years data have been presented that relate NADPH oxidases and autophagy (177). Expression of GRP78, an ER chaperone and a required component of autophagy in mammalian cells (178), is prevented by Nox4 siRNA treatment during ER stress in human VSMCs (172). Moreover, Nox4-derived H2O2 has been shown to activate Ras/ERK in endothelial cells, leading to induction of autophagy, prevention cell death and enhanced differentiation (171). Some studies have also related Nox2 to Toll-like receptor-activated autophagy in phagocytes (179).

9. HYPOXIA

A primary function of blood vessels is to provide nutrients and O2 to maintain viability and physiological functions. Maintenance of oxygen homeostasis is crucial for development and evolution (180). Chronic hypoxia is considered to be a major mechanism involved in aging as well chronic diseases like pulmonary arterial hypertension (PAH) and diabetes. During hypoxia (reduced O2 availability) adaptive mechanisms are induced in the vasculature. Hypoxia-inducible transcriptional (HIF) factors are induced that regulate genes involved in energy metabolism, angiogenesis and apoptosis (181, 182). One such factor, HIF-1alpha, is a major member of the HIF family and is normally degraded by the proteasome, but is stabilized in the presence of hypoxia (183). Interestingly, data in literature suggest that signaling induced by HIF is different among cell types. During development or regeneration processes, O2 is required due to cellular proliferation and hypertrophy. In such conditions O2 availability decreases and HIF-1 is activated. In response to tissue ischemia, vascular endothelial growth factor (VEGF), stem cell factor and PDGF are induced in a HIF-1-dependent manner (184, 185). These factors then promote recovery after hypoxic injury, as shown in the impaired recovery of function after femoral artery ligation in heterozygous HIF-1alpha mice (185).

ROS have an important role in O2 homeostasis in many studies and Nox family proteins are suggested to be involved. Nox2 appears to be an O2 sensor in pulmonary airway chemoreceptors (186); however Nox2 is not required for the hypoxic pulmonary vasoconstriction response (187). Another study proposed that Nox1 is involved in induction of HIF-1, because Nox1-transfected cells show increased expression of HIF-1 that is inhibited by catalase and or DPI (188). In contrast, Nox4 is proposed to be a target of HIF-1 since overexpression of HIF-1 increases Nox4 expression and its downregulation prevents such an increase (189). Mutation of a putative hypoxia-responsive element in the Nox4 promoter abolishes its induction by HIF-1. Nox4 has also been suggested to act as an oxygen sensor in conjunction with the TWIK-related acid-sensitive K+ channel TASK-1 (190, 191).

10. PULMONARY HYPERTENSION

Chronic PAH is an important clinical condition present in wide range of cardiovascular patients. The development of pulmonary hypertension is characterized by chronic hypoxia, endothelial injury and subsequent inflammation (192). The pulmonary circulation responds to local airway hypoxia with vasoconstriction in order to direct blood flow to better-ventilated regions of the lung. However, chronic hypoxia causes vascular remodeling associated with a loss of relaxation to endogenous NO, increased levels of O2·− and consequently increased pulmonary arterial pressure (193). As expected, ROS are involved in acute hypoxic vasoconstriction in the lung, as demonstrated by attenuation of vasoconstriction in the presence of SOD (194). Overexpression of extracellular SOD also attenuates vascular remodeling in chronic hypoxia (195).

In terms of the source of ROS involved in this response, many investigators have reported a role for NADPH oxidases. For example, Nox1 is upregulated in pulmonary resistance arteries of newborn piglets exposed to long-term hypoxia (196). Nox2-deficient mice show impairment of the vasoconstrictor response to ET-1 in the pulmonary vasculature after exposure to chronic hypoxia (197). Moreover, p47phox-deficient mice exhibit a reduction of acute hypoxic pulmonary vasoconstriction, while non-hypoxia-induced vasoconstrictions are not affected (198). Disruption of the Nox2 gene also has chronic effects: it inhibits hypoxia-induced PAH and vascular remodeling (199, 200). A number of other studies, however, suggest that it is Nox4 that has an important role in the vascular remodeling process. Nox4 expression, but not that of the other Nox proteins, is upregulated in pulmonary artery smooth muscle cells (PASMC) of mice subjected to hypoxic conditions (201). This upregulation occurs via autocrine production of TGFbeta and insulin-like growth factor binding protein-3, and is necessary for PASMC proliferation (98). Another study found that downregulation of Nox4, but not of Nox2, promotes attenuation of hypoxic pulmonary vasoconstriction in bovine pulmonary arteries (202). The mechanisms by which hypoxia induces Nox4 and by which Nox4 influences vasoconstriction and remodeling are somewhat unclear, but likely involve the HIF family proteins and proliferator-activated receptor-gamma (PPARgamma). ROS-derived from Nox4 increase the activity of HIF-1 (203) and HIF-2alpha in PASMC (204). Diebold et al (189) found that hypoxia rapidly increases Nox4 mRNA and protein levels in PASMC as well as pulmonary vessels from mice in a HIF-1 dependent manner. PPARgamma appears to regulate hypoxia-stimulated Nox4 induction through an NFkappaB-mediated mechanism in human PASMC (205). It is clear that the contribution of NADPH oxidases to pulmonary hypertension is complex, and that more work is needed to fully understand the mechanisms involved.

11. ATHEROSCLEROSIS

Atherosclerosis is a long-term inflammatory disease considered to be a public health issue worldwide. The disease consists of formation of atherosclerotic plaques in hemodynamically unstable regions of the vessel. Risk factors associated with disease development and progression are diverse, but include genetic propensity, obesity, hypertension, hypercholesterolemia, diabetes and smoking. Many of the steps of the progression of disease are ROS-sensitive. Plaque development depends on the subendothelial retention of lipids; cytokines that induce inflammatory responses including endothelial activation and recruitment of immune cells (macrophages, dendritic cells, neutrophils, B and T cells); cell proliferation, migration, and secretion of extracellular matrix proteins; and cell death and activation of coagulation factors (206). All these processes have all been associated with activation of Nox family proteins.

Early in disease development, macrophages and other cells of the immune system that are rich in Nox2 invade the vessel wall, and induction and activation of other Nox proteins occur in resident vascular cells (39). ROS-derived from Nox enzymes contribute to macrophage-mediated oxidation of LDL, which can then activate Nox proteins in vascular cells, resulting in expression of adhesion molecules and the recruitment of immune cells (207, 208). It has been shown that p47phox is required for atherosclerotic plaque evolution in ApoE transgenic mice (209, 210). Other data also support a role for Nox family proteins in atherosclerosis, such as the observation that Nox1 mRNA and Nox4 mRNA in nonphagocytic cells, as well as expression of p22phox in phagocytes, are increased in atherosclerotic human coronary arteries (39, 130, 211, 212). In addition, recent work showed that knockout of Nox1 in mice lacking the ApoE gene also reduces lesion formation (211).

In the last few years, the relevance of ER stress and autophagy to the progression of atherosclerotic plaques has begun to be appreciated (213215). Human atherosclerotic plaques show increased ER stress as assessed by expression of GRP78 and the pro-apoptotic factor CHOP. In the same study, the authors found 7-Ketocholesterol (cited above as an inducer of Nox4) in the fibrous caps of atherosclerotic plaques. Treatment of cultured coronary artery smooth muscle cells with 7- Ketocholesterol induced the upregulation of chaperones and apoptosis, and this effect was abolished by antioxidants (172). Apoptosis of macrophages regulated by CHOP, and as a consequence an increase of the instability atherosclerotic plaques, has also been described by different groups (172, 216). More specifically, Thorp et al (217) tested the effect of CHOP deficiency in vivo by measuring aortic root lesions of fat-fed Chop+/+ApoE−/− and Chop−/− ApoE−/− mice. They found that plaque necrosis was reduced by approximately 50% and lesional apoptosis by 35% in the CHOP-deficient mice.

As part of the evolution of atherosclerosis, chronic inflammatory processes become associated with Toll-like receptors (TLRs), regulators of innate immunity (218). It has been demonstrated that TLRs can promote activation of Nox proteins. Stimulation of TLR4 with LPS induces ROS generation and NFkappaB activation in HEK293 cells via a mechanism involving Nox4 (219). Nox4-dependent ROS generation is also important for generation of proinflamatory cytokines by LPS in endothelial cells (220). Recently, palmitate was proposed to induce inflammation in human endothelial cells. Like LPS, palmitate increases O2·− production and stimulates NFkappaB signaling via TLR4-mediated activation of Nox4 (221). Finally, high-fat feeding increases expression of Nox4 and bone morphogenic protein (BMP4) in thoracic aorta in wild type, but not TLR4−/− mice (221).

12. ANGIOGENESIS

Angiogenesis consists of the formation of new capillaries from preexisting vessels and is part of the process of neovascularization. During angiogenesis, endothelial cells proliferate, migrate and form capillary tubes. As described above, ROS play a crucial role in all of these processes. Importantly, Nox family proteins are activated by angiogenic stimuli such as VEGF, Ang II, hypoxia, cytokines and shear stress (34, 222). Transgenic mice overexpressing p22phox have increased VEGF expression and enhanced angiogenesis in experimental atheromas (223). Rac-1 mediated O2·− production mediates VEGF-induced placental angiogenesis (224) as well as that induced by overexpression of phosphodiesterase-2 (225). More specifically, a wide range of studies suggests a role for Nox2 in angiogenesis (226, 227). Studies using the hindlimb ischemia model in mice showed that Nox2 expression and ROS production are increased in bone marrow mononuclear cells (228). However, mice lacking Nox2 have impaired ischemia-induced blood flow recovery and neovascularization, which is associated with reductionof ROS production in bone marrow-derived cells as well as with a decrease in the number of endothelial progenitor cells in peripheral blood (228). Another study exposed p47phox deficient mice to arsenite in the drinking water. Environmental arsenic is a risk factor for developing vascular disease marked by pathologic remodeling. p47phox deficient mice were protected from arsenite-induced capillarization. Ex vivo arsenic exposure of endothelial cells led to increased O2·− generation that was inhibited by downregulation of Nox2, suggesting that Nox2 is required for capillarization (229).

Nox2 is not the only Nox involved, however: Nox1 overexpression can induce expression of VEGF and its receptors and upregulate matrix metalloproteinase activity through ROS generation, thus promoting angiogenesis (230). Nox1 downregulation decreases endothelial migration and tube-like structure formation by a mechanism involving NFkappaB signaling (137). Notably, mice deficient in Nox1, but not Nox4, exhibit impaired tumor angiogenesis (231). The role of Nox4 may be context dependent, however. Nox4-null animals exposed to chronic pressure overload develop contractile dysfunction, hypertrophy, and cardiac dilatation, whereas Nox4-trangenic mice are protected. The mechanism appears to be related to Nox4-dependent preservation of myocardial capillary density after pressure overload. Overexpression of Nox5 induces endothelial proliferation and formation of capillary-like structures (22).

Alternatively, excess ROS production can impair angiogenesis. An example of this involves the effects of Ang II in formation of new vessels after ischemia. Ang II signaling has different effects on angiogenesis depending on the basal vascular oxidative stress. Elevated levels of basal ROS make the effect of Ang II on collateral growth inhibitory, while Ang II administered when basal levels of ROS are low promotes collateral growth (232). Another case where excess ROS have been shown to be inhibitory is in pulmonary artery endothelial cells from fetal lambs with persistent pulmonary hypertension. These cells have increased expression of Nox subunits (Nox4, Nox2, Rac1, p47phox and p67phox) and increased Nox activity, but exhibit decreased tube formation, cell proliferation, scratch recovery and cell invasion, suggesting that increased oxidative stress from Nox activity contributes to impaired angiogenesis in this model (233).

13. DIABETIC VASCULAR DISEASE

Diabetes mellitus is accompanied by an increased risk of developing cardiovascular disease. A wide range of studies has implicated oxidative stress in the pathology of diabetes. Elevated oxidative stress is detected in diabetic patients and in animal models of diabetes (obese ob/ob mice, Zucker fatty rats, streptozotocin-induced diabetes) (234).

Hyperglycemia, hyperinsulinemia, and increased levels of lipids promote ROS generation in diverse cell types. The enzymatic sources of ROS include Nox family proteins, mitochondria and uncoupled NO synthases, with a very strong probability of cross-talk between them. Enhanced expression of p22phox, increased Nox activity and elevated O2·− production were found in mouse microvascular endothelial cells exposed to high glucose (235). High glucose, advanced glycation endproducts and free fatty acids induce expression and activity of Nox in VSMCs as well (236239). In one study, downregulation of p47phox in VSMCs decreased O2·− generation (235), and no apparent role for eNOS, mitochondria or xanthine oxidase was found (240). In contrast, another group showed that ROS generation by hyperglycemia was decreased by an inhibitor of electron transport chain complex II, uncoupling protein-1 or MnSOD in aortic endothelial cells (241). In aortas of streptozotocin diabetic rats, uncoupled eNOS and Nox2 have been implicated in ROS generation (242). It is thus clear that there are multiple sources of ROS in the diabetic vasculature, depending on cell type, stimulus and diabetic model.

PKC appears to be extremely important in Nox regulation and activation in diabetes. For example, activation of Nox was abolished in diabetic PKCbeta knockout mice (243). Other isoforms of PKC have been implicated as well: PKCalpha mediates Nox activation in the kidney of diabetic rats, PKCdelta in adipocytes (244) and PKCzeta in mesangial cells treated with high glucose (245). JNK activation and Akt inhibition may be downstream of Nox2-derived ROS, at least in beta-cells exposed to high glucose (246).

Insulin increases ROS generation in VSMCs, adipocytes, fibroblasts and macrophages (247250). In macrophages, ROS induced by insulin are derived from Nox2 (250). In other insulin-sensitive cells, Nox4 appears to have important role in the response to insulin since its downregulation prevents ROS production and inactivation of phosphatase PTP1B, a widely expressed phosphatase implicated in the negative regulation of insulin signaling (251). Another study suggests that the upregulation of Nox4 by TGFbeta in VSMCs from obese Zucker rats is responsible for the impaired response to NO by a mechanism involving the oxidation of SERCA C674. Such mechanisms could explain the enhanced VSMC migration and neointima formation in type 2 diabetes (252).

Another effect of excess ROS generation in diabetes is impairment of acetylcholine-induced endothelium-dependent relaxation, a response that has been attributed to NADPH oxidases, at least in OLETF rats (253). Downregulation of Nox enzymes has also been shown to decrease the proinflammatory effects of AGEs in different cell types. For example, macrophages from Nox2−/− mice showed inhibition of AGE effects (254). Downregulation of Nox1 impairs AGE-induced ROS production and induction of iNOS in VSMCS (237). Finally, treatment of db/db mice (a model of type 2 diabetes) with the O2·− scavenger tempol reduces expression of inflammatory genes in the vessel wall (236).

Pathways that inhibit Nox-dependent ROS production in diabetes are thus quite promising therapeutic targets for diabetic vascular dysfunction. One mechanism that reduces the enhanced PDGF-induced ROS generation that occurs in response to high glucose in VSMC is activation of the protein tyrosine phosphatase PTPepsilonM. PTPepsilonM negatively regulates PDGFbeta-receptor signaling, DNA synthesis and migration by reducing the phosphorylation level of the PDGFbeta receptor with subsequent downregulation of Nox1 and suppression of H2O2 generation (255). Another inhibitor of diabetes-induced oxidative stress is resveratrol. Resveratrol is a natural phytophenol found in wine that exhibits cardioprotective effects. It has been shown to reduce oxidative/nitrative stress and improve NO availability by inhibition of TNFalpha-induced NF-kappaB activation and expression and activation of Nox2 and iNOS in type 2 diabetes (256).

14. PERSPECTIVE

Studies summarized in this review make it clear that Nox-derived ROS play a role in virtually all aspects of vascular function, both physiologically and pathophysiologically (Figure 1). Individual Nox enzymes play distinct roles both within a cell and between cell types. Moreover, Nox proteins exhibit considerable cross-talk with other oxidant-generating systems (33). As we begin to understand this differential regulation of ROS-dependent signaling, we will be able to more rationally develop therapies targeted to specific sources of ROS, rather than relying on global antioxidant treatments that have been suboptimal in the past. The complexities of redox regulation of vascular function provide fertile ground for future investigation.

Figure 1.

Figure 1

Diverse roles of NADPH oxidases in vascular physiology and pathophysiology. ROS generation by Nox enzymes maintain vascular tone and integrity, participate in the response to injury, and contribute to vascular diseases.

Acknowledgments

This work was supported by NIH grants HL38206, HL092120, HL095070 and HL058863; and AHA 09POST2070000.

The abbreviations used are

Ang II

angiotensin II

ApoE

Apolipoprotein E

ASK1

apoptosis signal-regulating kinase

bFGF

basic fibroblast growth factor

BH4

tetrahydrobiopterin

CaMKII

Ca2+/calmodulin-dependent protein kinase II

CHOP

C/EBP-homologous protein

CO

carbon monoxide

ECM

extracellular matrix

EDHF

endothelium-dependent hyperpolarizing factor

eNOS

endothelial nitric oxide synthase

ER

endoplasmic reticulum

ERK1/2

extracellular signal regulated kinase

ERO-1

ER oxidoreductin 1

ET-1

endothelin-1

H2O2

hydrogen peroxide

HIF

Hypoxia-inducible factor

HO-1

heme oxygenase

HSP

heat shock protein

IGF-1

insulin-like growth factor-1

IKKalpha

inhibitor of NFkappaB kinase-alpha

JNK

c-Jun N-terminal kinase

LPS

lipopolysaccharide

MAPK

mitogen-activated protein kinase

MKP-1

MAPK phosphatase-1

MMP

matrix metalloproteinase

MnSOD

manganese superoxide dismutase

MRTF

myocardin-related transcription factor

NFkappaB

nuclear factor kappa B

NO

nitric oxide

NOX

NADPH oxidase

O2−·

superoxide anion

OLETF

Otsuka Long-Evans Tokushima Fatty

PAH

pulmonary arterial hypertension

PASMC

pulmonary artery smooth muscle cells

PDGF

platelet-derived growth factor

PDI

protein disulfide isomerase

PI3K

phosphoinositide-3 kinase

PKC

protein kinase C

PPARgama

peroxisome proliferator-activated receptor-gamma

PTP

tyrosine-protein phosphatase

ROS

reactive oxygen species

SERCA

sarco/endoplasmic reticulum ATPase

SHP2

SH2 domain–containing protein tyrosine phosphatase–2

SRF

serum-response factor

SSH1L

slingshot-1 phosphatase

TASK1

Tandem-pore acid-sensing potassium channel-1

TGFbeta

transforming growth factor beta

TLR

toll-like receptors

TNFalpha

tumor necrosis factor alpha

TRAF2

TNF receptor-associated factor 2

UPR

unfolded protein response

VEGF

vascular endothelial growth factor

VSMC

vascular smooth muscle cells

References

  • 1.D’Autreaux B, Toledano MB. ROS as signalling molecules: mechanisms that generate specificity in ROS homeostasis. Nat Rev Mol Cell Biol. 2007;8(10):813–24. doi: 10.1038/nrm2256. [DOI] [PubMed] [Google Scholar]
  • 2.Lambeth JD. Nox enzymes, ROS, and chronic disease: an example of antagonistic pleiotropy. Free Radic Biol Med. 2007;43(3):332–47. doi: 10.1016/j.freeradbiomed.2007.03.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Mueller CF, Laude K, McNally JS, Harrison DG. ATVB in focus: redox mechanisms in blood vessels. Arterioscler Thromb Vasc Biol. 2005;25(2):274–8. doi: 10.1161/01.ATV.0000149143.04821.eb. [DOI] [PubMed] [Google Scholar]
  • 4.Forstermann U. Oxidative stress in vascular disease: causes, defense mechanisms and potential therapies. Nat Clin Pract Cardiovasc Med. 2008;5(6):338–49. doi: 10.1038/ncpcardio1211. [DOI] [PubMed] [Google Scholar]
  • 5.Go YM, Jones DP. Cysteine/cystine redox signaling in cardiovascular disease. Free Radic Biol Med. 2011;50(4):495–509. doi: 10.1016/j.freeradbiomed.2010.11.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Go YM, Jones DP. Redox control systems in the nucleus: mechanisms and functions. Antioxid Redox Signal. 2010;13(4):489–509. doi: 10.1089/ars.2009.3021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Go YM, Jones DP. Redox compartmentalization in eukaryotic cells. Biochim Biophys Acta. 2008;1780(11):1273–90. doi: 10.1016/j.bbagen.2008.01.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Lassegue B, Griendling KK. NADPH oxidases: functions and pathologies in the vasculature. Arterioscler Thromb Vasc Biol. 2010;30(4):653–61. doi: 10.1161/ATVBAHA.108.181610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Lyle AN, Griendling KK. Modulation of vascular smooth muscle signaling by reactive oxygen species. Physiology (Bethesda) 2006;21:269–80. doi: 10.1152/physiol.00004.2006. [DOI] [PubMed] [Google Scholar]
  • 10.Brandes RP, Schroder K. Composition and functions of vascular nicotinamide adenine dinucleotide phosphate oxidases. Trends Cardiovasc Med. 2008;18(1):15–9. doi: 10.1016/j.tcm.2007.11.001. [DOI] [PubMed] [Google Scholar]
  • 11.Gorlach A, Klappa P, Kietzmann T. The endoplasmic reticulum: folding, calcium homeostasis, signaling, and redox control. Antioxid Redox Signal. 2006;8(9–10):1391–418. doi: 10.1089/ars.2006.8.1391. [DOI] [PubMed] [Google Scholar]
  • 12.Taylor CT, Moncada S. Nitric oxide, cytochrome C oxidase, and the cellular response to hypoxia. Arterioscler Thromb Vasc Biol. 2010;30(4):643–7. doi: 10.1161/ATVBAHA.108.181628. [DOI] [PubMed] [Google Scholar]
  • 13.Jones DP. Redefining oxidative stress. Antioxid Redox Signal. 2006;8(9–10):1865–79. doi: 10.1089/ars.2006.8.1865. [DOI] [PubMed] [Google Scholar]
  • 14.Brown DI, Griendling KK. Nox proteins in signal transduction. Free Radic Biol Med. 2009;47(9):1239–53. doi: 10.1016/j.freeradbiomed.2009.07.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Babior BM, Lambeth JD, Nauseef W. The neutrophil NADPH oxidase. Arch Biochem Biophys. 2002;397(2):342–4. doi: 10.1006/abbi.2001.2642. [DOI] [PubMed] [Google Scholar]
  • 16.Suh YA, Arnold RS, Lassegue B, Shi J, Xu X, Sorescu D, Chung AB, Griendling KK, Lambeth JD. Cell transformation by the superoxide-generating oxidase Mox1. Nature. 1999;401(6748):79–82. doi: 10.1038/43459. [DOI] [PubMed] [Google Scholar]
  • 17.Banfi B, Maturana A, Jaconi S, Arnaudeau S, Laforge T, Sinha B, Ligeti E, Demaurex N, Krause KH. A mammalian H+ channel generated through alternative splicing of the NADPH oxidase homolog NOH-1. Science. 2000;287(5450):138–42. doi: 10.1126/science.287.5450.138. [DOI] [PubMed] [Google Scholar]
  • 18.Lambeth JD, Kawahara T, Diebold B. Regulation of Nox and Duox enzymatic activity and expression. Free Radic Biol Med. 2007;43(3):319–31. doi: 10.1016/j.freeradbiomed.2007.03.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Kawahara T, Quinn MT, Lambeth JD. Molecular evolution of the reactive oxygen-generating NADPH oxidase (Nox/Duox) family of enzymes. BMC Evol Biol. 2007;7:109. doi: 10.1186/1471-2148-7-109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Cheng G, Cao Z, Xu X, van Meir EG, Lambeth JD. Homologs of gp91phox: cloning and tissue expression of Nox3, Nox4, and Nox5. Gene. 2001;269(1–2):131–40. doi: 10.1016/s0378-1119(01)00449-8. [DOI] [PubMed] [Google Scholar]
  • 21.Banfi B, Molnar G, Maturana A, Steger K, Hegedus B, Demaurex N, Krause KH. A Ca (2+)-activated NADPH oxidase in testis, spleen, and lymph nodes. J Biol Chem. 2001;276(40):37594–601. doi: 10.1074/jbc.M103034200. [DOI] [PubMed] [Google Scholar]
  • 22.BelAiba RS, Djordjevic T, Petry A, Diemer K, Bonello S, Banfi B, Hess J, Pogrebniak A, Bickel C, Gorlach A. NOX5 variants are functionally active in endothelial cells. Free Radic Biol Med. 2007;42(4):446–59. doi: 10.1016/j.freeradbiomed.2006.10.054. [DOI] [PubMed] [Google Scholar]
  • 23.Geiszt M, Lekstrom K, Witta J, Leto TL. Proteins homologous to p47phox and p67phox support superoxide production by NAD (P)H oxidase 1 in colon epithelial cells. J Biol Chem. 2003;278(22):20006–12. doi: 10.1074/jbc.M301289200. [DOI] [PubMed] [Google Scholar]
  • 24.Lyle AN, Deshpande NN, Taniyama Y, Seidel-Rogol B, Pounkova L, Du P, Papaharalambus C, Lassegue B, Griendling KK. Poldip2, a novel regulator of Nox4 and cytoskeletal integrity in vascular smooth muscle cells. Circ Res. 2009;105(3):249–59. doi: 10.1161/CIRCRESAHA.109.193722. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Banfi B, Clark RA, Steger K, Krause KH. Two novel proteins activate superoxide generation by the NADPH oxidase NOX1. J Biol Chem. 2003;278(6):3510–3. doi: 10.1074/jbc.C200613200. [DOI] [PubMed] [Google Scholar]
  • 26.Takeya R, Ueno N, Kami K, Taura M, Kohjima M, Izaki T, Nunoi H, Sumimoto H. Novel human homologues of p47phox and p67phox participate in activation of superoxide-producing NADPH oxidases. J Biol Chem. 2003;278(27):25234–46. doi: 10.1074/jbc.M212856200. [DOI] [PubMed] [Google Scholar]
  • 27.Ueyama T, Lekstrom K, Tsujibe S, Saito N, Leto TL. Subcellular localization and function of alternatively spliced Noxo1 isoforms. Free Radic Biol Med. 2007;42(2):180–90. doi: 10.1016/j.freeradbiomed.2006.08.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Viedt C, Fei J, Krieger-Brauer HI, Brandes RP, Teupser D, Kamimura M, Katus HA, Kreuzer J. Role of p22phox in angiotensin II and platelet-derived growth factor AA induced activator protein 1 activation in vascular smooth muscle cells. J Mol Med. 2004;82(1):31–8. doi: 10.1007/s00109-003-0500-5. [DOI] [PubMed] [Google Scholar]
  • 29.Ambasta RK, Kumar P, Griendling KK, Schmidt HH, Busse R, Brandes RP. 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(44):45935–41. doi: 10.1074/jbc.M406486200. [DOI] [PubMed] [Google Scholar]
  • 30.Hanna IR, Hilenski LL, Dikalova A, Taniyama Y, Dikalov S, Lyle A, Quinn MT, Lassegue B, Griendling KK. Functional association of nox1 with p22phox in vascular smooth muscle cells. Free Radic Biol Med. 2004;37(10):1542–9. doi: 10.1016/j.freeradbiomed.2004.08.011. [DOI] [PubMed] [Google Scholar]
  • 31.Nauseef WM. Biological roles for the NOX family NADPH oxidases. J Biol Chem. 2008;283(25):16961–5. doi: 10.1074/jbc.R700045200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Brandes RP, Weissmann N, Schroder K. NADPH oxidases in cardiovascular disease. Free Radic Biol Med. 2010;49(5):687–706. doi: 10.1016/j.freeradbiomed.2010.04.030. [DOI] [PubMed] [Google Scholar]
  • 33.Datla SR, Griendling KK. Reactive oxygen species, NADPH oxidases, and hypertension. Hypertension. 2010;56(3):325–30. doi: 10.1161/HYPERTENSIONAHA.109.142422. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Guzik TJ, Griendling KK. NADPH oxidases: molecular understanding finally reaching the clinical level? Antioxid Redox Signal. 2009;11(10):2365–70. doi: 10.1089/ars.2009.2615. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Williams HC, Griendling KK. NADPH oxidase inhibitors: new antihypertensive agents? J Cardiovasc Pharmacol. 2007;50(1):9–16. doi: 10.1097/FJC.0b013e318063e820. [DOI] [PubMed] [Google Scholar]
  • 36.Brandes RP. Out of balance: a role of impaired superoxide dismutase activity for vascular constrictive remodeling after angioplasty. Arterioscler Thromb Vasc Biol. 2003;23(12):2121–2. doi: 10.1161/01.ATV.0000099269.04527.88. [DOI] [PubMed] [Google Scholar]
  • 37.Hilenski LL, Clempus RE, Quinn MT, Lambeth JD, Griendling KK. Distinct subcellular localizations of Nox1 and Nox4 in vascular smooth muscle cells. Arterioscler Thromb Vasc Biol. 2004;24(4):677–83. doi: 10.1161/01.ATV.0000112024.13727.2c. [DOI] [PubMed] [Google Scholar]
  • 38.Helmcke I, Heumuller S, Tikkanen R, Schroder K, Brandes RP. Identification of structural elements in Nox1 and Nox4 controlling localization and activity. Antioxid Redox Signal. 2009;11(6):1279–87. doi: 10.1089/ars.2008.2383. [DOI] [PubMed] [Google Scholar]
  • 39.Lassegue B, Griendling KK. NADPH oxidases: functions and pathologies in the vasculature. Arterioscler Thromb Vasc Biol. 2010;30(4):653–61. doi: 10.1161/ATVBAHA.108.181610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Serrander L, Cartier L, Bedard K, Banfi B, Lardy B, Plastre O, Sienkiewicz A, Forro L, Schlegel W, Krause KH. NOX4 activity is determined by mRNA levels and reveals a unique pattern of ROS generation. Biochem J. 2007;406(1):105–14. doi: 10.1042/BJ20061903. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Janiszewski M, Lopes LR, Carmo AO, Pedro MA, Brandes RP, Santos CX, Laurindo FR. Regulation of NAD (P)H oxidase by associated protein disulfide isomerase in vascular smooth muscle cells. J Biol Chem. 2005;280(49):40813–9. doi: 10.1074/jbc.M509255200. [DOI] [PubMed] [Google Scholar]
  • 42.Lassegue B, Sorescu D, Szocs K, Yin Q, Akers M, Zhang Y, Grant SL, Lambeth JD, Griendling KK. 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(9):888–94. doi: 10.1161/hh0901.090299. [DOI] [PubMed] [Google Scholar]
  • 43.Santos CX, Tanaka LY, Wosniak J, Laurindo FR. Mechanisms and implications of reactive oxygen species generation during the unfolded protein response: roles of endoplasmic reticulum oxidoreductases, mitochondrial electron transport, and NADPH oxidase. Antioxid Redox Signal. 2009;11(10):2409–27. doi: 10.1089/ars.2009.2625. [DOI] [PubMed] [Google Scholar]
  • 44.Laurindo FR, Fernandes DC, Amanso AM, Lopes LR, Santos CX. Novel role of protein disulfide isomerase in the regulation of NADPH oxidase activity: pathophysiological implications in vascular diseases. Antioxid Redox Signal. 2008;10(6):1101–13. doi: 10.1089/ars.2007.2011. [DOI] [PubMed] [Google Scholar]
  • 45.Santos CX, Stolf BS, Takemoto PV, Amanso AM, Lopes LR, Souza EB, Goto H, Laurindo FR. Protein disulfide isomerase (PDI) associates with NADPH oxidase and is required for phagocytosis of Leishmania chagasi promastigotes by macrophages. J Leukoc Biol. 2009;86(4):989–98. doi: 10.1189/jlb.0608354. [DOI] [PubMed] [Google Scholar]
  • 46.Jay DB, Papaharalambus CA, Seidel-Rogol B, Dikalova AE, Lassegue B, Griendling KK. Nox5 mediates PDGF-induced proliferation in human aortic smooth muscle cells. Free Radic Biol Med. 2008;45(3):329–35. doi: 10.1016/j.freeradbiomed.2008.04.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Baysal BE. A phenotypic perspective on Mammalian oxygen sensor candidates. Ann N Y Acad Sci. 2006;1073:221–33. doi: 10.1196/annals.1353.024. [DOI] [PubMed] [Google Scholar]
  • 48.Palmer RM, Ashton DS, Moncada S. Vascular endothelial cells synthesize nitric oxide from L-arginine. Nature. 1988;333(6174):664–6. doi: 10.1038/333664a0. [DOI] [PubMed] [Google Scholar]
  • 49.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(12):1225–32. doi: 10.1161/01.RES.0000104087.29395.66. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Price DT, Vita JA, Keaney JF., Jr Redox control of vascular nitric oxide bioavailability. Antioxid Redox Signal. 2000;2(4):919–35. doi: 10.1089/ars.2000.2.4-919. [DOI] [PubMed] [Google Scholar]
  • 51.Halliwell B, Zhao K, Whiteman M. Nitric oxide peroxynitrite. The ugly, the uglier and the not so good: a personal view of recent controversies. Free Radic Res. 1999;31(6):651–69. doi: 10.1080/10715769900301221. [DOI] [PubMed] [Google Scholar]
  • 52.Jeremy JY, Yim AP, Wan S, Angelini GD. Oxidative stress, nitric oxide, and vascular disease. J Card Surg. 2002;17(4):324–7. doi: 10.1111/j.1540-8191.2001.tb01151.x. [DOI] [PubMed] [Google Scholar]
  • 53.Zou MH, Shi C, Cohen RA. Oxidation of the zinc-thiolate complex and uncoupling of endothelial nitric oxide synthase by peroxynitrite. J Clin Invest. 2002;109(6):817–26. doi: 10.1172/JCI14442. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Zou MH. Peroxynitrite and protein tyrosine nitration of prostacyclin synthase. Prostaglandins Other Lipid Mediat. 2007;82(1–4):119–27. doi: 10.1016/j.prostaglandins.2006.05.005. [DOI] [PubMed] [Google Scholar]
  • 55.Fujimoto S, Asano T, Sakai M, Sakurai K, Takagi D, Yoshimoto N, Itoh T. Mechanisms of hydrogen peroxide-induced relaxation in rabbit mesenteric small artery. Eur J Pharmacol. 2001;412(3):291–300. doi: 10.1016/s0014-2999(00)00940-7. [DOI] [PubMed] [Google Scholar]
  • 56.Gao YJ, Hirota S, Zhang DW, Janssen LJ, Lee RM. Mechanisms of hydrogen-peroxide-induced biphasic response in rat mesenteric artery. Br J Pharmacol. 2003;138(6):1085–92. doi: 10.1038/sj.bjp.0705147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Burke TM, Wolin MS. Hydrogen peroxide elicits pulmonary arterial relaxation and guanylate cyclase activation. Am J Physiol. 1987;252(4 Pt 2):H721–32. doi: 10.1152/ajpheart.1987.252.4.H721. [DOI] [PubMed] [Google Scholar]
  • 58.Lucchesi PA, Belmadani S, Matrougui K. Hydrogen peroxide acts as both vasodilator and vasoconstrictor in the control of perfused mouse mesenteric resistance arteries. J Hypertens. 2005;23(3):571–9. doi: 10.1097/01.hjh.0000160214.40855.79. [DOI] [PubMed] [Google Scholar]
  • 59.Liu Y, Bubolz AH, Mendoza S, Zhang DX, Gutterman DD. H2O2 is the transferrable factor mediating flow-induced dilation in human coronary arterioles. Circ Res. 2011;108(5):566–73. doi: 10.1161/CIRCRESAHA.110.237636. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Matoba T, Shimokawa H, Kubota H, Morikawa K, Fujiki T, Kunihiro I, Mukai Y, Hirakawa Y, Takeshita A. Hydrogen peroxide is an endothelium-derived hyperpolarizing factor in human mesenteric arteries. Biochem Biophys Res Commun. 2002;290(3):909–13. doi: 10.1006/bbrc.2001.6278. [DOI] [PubMed] [Google Scholar]
  • 61.Matoba T, Shimokawa H, Nakashima M, Hirakawa Y, Mukai Y, Hirano K, Kanaide H, Takeshita A. Hydrogen peroxide is an endothelium-derived hyperpolarizing factor in mice. J Clin Invest. 2000;106(12):1521–30. doi: 10.1172/JCI10506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Suvorava T, Lauer N, Kumpf S, Jacob R, Meyer W, Kojda G. Endogenous vascular hydrogen peroxide regulates arteriolar tension in vivo. Circulation. 2005;112(16):2487–95. doi: 10.1161/CIRCULATIONAHA.105.543157. [DOI] [PubMed] [Google Scholar]
  • 63.Ellis A, Pannirselvam M, Anderson TJ, Triggle CR. Catalase has negligible inhibitory effects on endothelium-dependent relaxations in mouse isolated aorta and small mesenteric artery. Br J Pharmacol. 2003;140(7):1193–200. doi: 10.1038/sj.bjp.0705549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Oeckler RA, Arcuino E, Ahmad M, Olson SC, Wolin MS. Cytosolic NADH redox and thiol oxidation regulate pulmonary arterial force through ERK MAP kinase. Am J Physiol Lung Cell Mol Physiol. 2005;288(6):L1017–25. doi: 10.1152/ajplung.00223.2004. [DOI] [PubMed] [Google Scholar]
  • 65.Oeckler RA, Kaminski PM, Wolin MS. Stretch enhances contraction of bovine coronary arteries via an NAD (P)H oxidase-mediated activation of the extracellular signal-regulated kinase mitogen-activated protein kinase cascade. Circ Res. 2003;92(1):23–31. doi: 10.1161/01.res.0000051860.84509.ce. [DOI] [PubMed] [Google Scholar]
  • 66.Knock GA, Snetkov VA, Shaifta Y, Connolly M, Drndarski S, Noah A, Pourmahram GE, Becker S, Aaronson PI, Ward JP. Superoxide constricts rat pulmonary arteries via Rho-kinase-mediated Ca (2+) sensitization. Free Radic Biol Med. 2009;46(5):633–42. doi: 10.1016/j.freeradbiomed.2008.11.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Larsen BT, Bubolz AH, Mendoza SA, Pritchard KA, Jr, Gutterman DD. Bradykinin-induced dilation of human coronary arterioles requires NADPH oxidase-derived reactive oxygen species. Arterioscler Thromb Vasc Biol. 2009;29(5):739–45. doi: 10.1161/ATVBAHA.108.169367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Schiffrin EL. Remodeling of resistance arteries in essential hypertension and effects of antihypertensive treatment. Am J Hypertens. 2004;17(12 Pt 1):1192–200. doi: 10.1016/j.amjhyper.2004.05.023. [DOI] [PubMed] [Google Scholar]
  • 69.Rizzoni D, Rodella L, Porteri E, Rezzani R, Guelfi D, Piccoli A, Castellano M, Muiesan ML, Bianchi R, Rosei EA. Time course of apoptosis in small resistance arteries of spontaneously hypertensive rats. J Hypertens. 2000;18(7):885–91. doi: 10.1097/00004872-200018070-00010. [DOI] [PubMed] [Google Scholar]
  • 70.Ward MR, Pasterkamp G, Yeung AC, Borst C. Arterial remodeling. Mechanisms and clinical implications. Circulation. 2000;102(10):1186–91. doi: 10.1161/01.cir.102.10.1186. [DOI] [PubMed] [Google Scholar]
  • 71.Hynes RO. The extracellular matrix: not just pretty fibrils. Science. 2009;326(5957):1216–9. doi: 10.1126/science.1176009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Duprez DA. Role of the renin-angiotensin-aldosterone system in vascular remodeling and inflammation: a clinical review. J Hypertens. 2006;24(6):983–91. doi: 10.1097/01.hjh.0000226182.60321.69. [DOI] [PubMed] [Google Scholar]
  • 73.Marchesi C, Ebrahimian T, Angulo O, Paradis P, Schiffrin EL. Endothelial nitric oxide synthase uncoupling and perivascular adipose oxidative stress and inflammation contribute to vascular dysfunction in a rodent model of metabolic syndrome. Hypertension. 2009;54(6):1384–92. doi: 10.1161/HYPERTENSIONAHA.109.138305. [DOI] [PubMed] [Google Scholar]
  • 74.Liu J, Yang F, Yang XP, Jankowski M, Pagano PJ. NAD (P)H oxidase mediates angiotensin II-induced vascular macrophage infiltration and medial hypertrophy. Arterioscler Thromb Vasc Biol. 2003;23(5):776–82. doi: 10.1161/01.ATV.0000066684.37829.16. [DOI] [PubMed] [Google Scholar]
  • 75.Liu J, Ormsby A, Oja-Tebbe N, Pagano PJ. Gene transfer of NAD (P)H oxidase inhibitor to the vascular adventitia attenuates medial smooth muscle hypertrophy. Circ Res. 2004;95(6):587–94. doi: 10.1161/01.RES.0000142317.88591.e6. [DOI] [PubMed] [Google Scholar]
  • 76.Dikalova A, Clempus R, Lassegue B, Cheng G, McCoy J, Dikalov S, San Martin A, Lyle A, Weber DS, Weiss D, Taylor WR, Schmidt HH, Owens GK, Lambeth JD, Griendling KK. Nox1 overexpression potentiates angiotensin II-induced hypertension and vascular smooth muscle hypertrophy in transgenic mice. Circulation. 2005;112(17):2668–76. doi: 10.1161/CIRCULATIONAHA.105.538934. [DOI] [PubMed] [Google Scholar]
  • 77.Matsuno K, Yamada H, Iwata K, Jin D, Katsuyama M, Matsuki M, Takai S, Yamanishi K, Miyazaki M, Matsubara H, Yabe-Nishimura C. Nox1 is involved in angiotensin II-mediated hypertension: a study in Nox1-deficient mice. Circulation. 2005;112(17):2677–85. doi: 10.1161/CIRCULATIONAHA.105.573709. [DOI] [PubMed] [Google Scholar]
  • 78.Janmey PA. The cytoskeleton and cell signaling: component localization and mechanical coupling. Physiol Rev. 1998;78(3):763–81. doi: 10.1152/physrev.1998.78.3.763. [DOI] [PubMed] [Google Scholar]
  • 79.Mancini A, Di Battista JA. Transcriptional regulation of matrix metalloprotease gene expression in health and disease. Front Biosci. 2006;11:423–46. doi: 10.2741/1809. [DOI] [PubMed] [Google Scholar]
  • 80.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 metalloproteinasesin vitro Implications for atherosclerotic plaque stability. J Clin Invest. 1996;98(11):2572–9. doi: 10.1172/JCI119076. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Gurjar MV, DeLeon J, Sharma RV, Bhalla RC. Mechanism of inhibition of matrix metalloproteinase-9 induction by NO in vascular smooth muscle cells. J Appl Physiol. 2001;91(3):1380–6. doi: 10.1152/jappl.2001.91.3.1380. [DOI] [PubMed] [Google Scholar]
  • 82.Browatzki M, Larsen D, Pfeiffer CA, Gehrke SG, Schmidt J, Kranzhofer A, Katus HA, Kranzhofer R. Angiotensin II stimulates matrix metalloproteinase secretion in human vascular smooth muscle cells via nuclear factor-kappaB and activator protein 1 in a redox-sensitive manner. J Vasc Res. 2005;42(5):415–23. doi: 10.1159/000087451. [DOI] [PubMed] [Google Scholar]
  • 83.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(11):e80–6. doi: 10.1161/01.RES.0000077044.60138.7C. [DOI] [PubMed] [Google Scholar]
  • 84.Chu X, Filali M, Stanic B, Takapoo M, Sheehan A, Bhalla R, Lamb FS, Miller FJ., Jr A critical role for chloride channel-3 (CIC-3) in smooth muscle cell activation and neointima formation. Arterioscler Thromb Vasc Biol. 2011;31(2):345–51. doi: 10.1161/ATVBAHA.110.217604. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Al Ghouleh I, Pagano PJ. Endosomal ClC-3 and Nox1: moving marksmen of redox signaling? Arterioscler Thromb Vasc Biol. 2011;31(2):240–2. doi: 10.1161/ATVBAHA.110.220053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Matsuda JJ, Filali MS, Moreland JG, Miller FJ, Lamb FS. Activation of swelling-activated chloride current by tumor necrosis factor-alpha requires ClC-3-dependent endosomal reactive oxygen production. J Biol Chem. 2010;285(30):22864–73. doi: 10.1074/jbc.M109.099838. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Hecker L, Vittal R, Jones T, Jagirdar R, Luckhardt TR, Horowitz JC, Pennathur S, Martinez FJ, Thannickal VJ. NADPH oxidase-4 mediates myofibroblast activation and fibrogenic responses to lung injury. Nat Med. 2009;15(9):1077–81. doi: 10.1038/nm.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Awad AE, Kandalam V, Chakrabarti S, Wang X, Penninger JM, Davidge ST, Oudit GY, Kassiri Z. Tumor necrosis factor induces matrix metalloproteinases in cardiomyocytes and cardiofibroblasts differentially via superoxide production in a PI3Kgamma-dependent manner. Am J Physiol Cell Physiol. 298(3):C679–92. doi: 10.1152/ajpcell.00351.2009. [DOI] [PubMed] [Google Scholar]
  • 89.Kim SY, Lee JG, Cho WS, Cho KH, Sakong J, Kim JR, Chin BR, Baek SH. Role of NADPH oxidase-2 in lipopolysaccharide-induced matrix metalloproteinase expression and cell migration. Immunol Cell Biol. 2010;88(2):197–204. doi: 10.1038/icb.2009.87. [DOI] [PubMed] [Google Scholar]
  • 90.Owens GK, Kumar MS, Wamhoff BR. Molecular regulation of vascular smooth muscle cell differentiation in development and disease. Physiol Rev. 2004;84(3):767–801. doi: 10.1152/physrev.00041.2003. [DOI] [PubMed] [Google Scholar]
  • 91.Finkel T. Oxidant signals and oxidative stress. Curr Opin Cell Biol. 2003;15(2):247–54. doi: 10.1016/s0955-0674(03)00002-4. [DOI] [PubMed] [Google Scholar]
  • 92.Finkel T. Reactive oxygen species and signal transduction. IUBMB Life. 2001;52(1–2):3–6. doi: 10.1080/15216540252774694. [DOI] [PubMed] [Google Scholar]
  • 93.Finkel T, Holbrook NJ. Oxidants, oxidative stress and the biology of ageing. Nature. 2000;408(6809):239–47. doi: 10.1038/35041687. [DOI] [PubMed] [Google Scholar]
  • 94.Sorescu D, Somers MJ, Lassegue B, Grant S, Harrison DG, Griendling KK. Electron spin resonance characterization of the NAD (P)H oxidase in vascular smooth muscle cells. Free Radic Biol Med. 2001;30(6):603–12. doi: 10.1016/s0891-5849(00)00507-4. [DOI] [PubMed] [Google Scholar]
  • 95.Deshpande NN, Sorescu D, Seshiah P, Ushio-Fukai M, Akers M, Yin Q, Griendling KK. Mechanism of hydrogen peroxide-induced cell cycle arrest in vascular smooth muscle. Antioxid Redox Signal. 2002;4(5):845–54. doi: 10.1089/152308602760599007. [DOI] [PubMed] [Google Scholar]
  • 96.Martin-Garrido A, Brown DI, Lyle AN, Dikalova A, Seidel-Rogol B, Lassegue B, San Martin A, Griendling KK. NADPH oxidase 4 mediates TGF-beta-induced smooth muscle alpha-actin via p38MAPK and serum response factor. Free Radic Biol Med. 2011;50(2):354–62. doi: 10.1016/j.freeradbiomed.2010.11.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Clempus RE, Sorescu D, Dikalova AE, Pounkova L, Jo P, Sorescu GP, Schmidt HH, Lassegue B, Griendling KK. Nox4 is required for maintenance of the differentiated vascular smooth muscle cell phenotype. Arterioscler Thromb Vasc Biol. 2007;27(1):42–8. doi: 10.1161/01.ATV.0000251500.94478.18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Ismail S, Sturrock A, Wu P, Cahill B, Norman K, Huecksteadt T, Sanders K, Kennedy T, Hoidal J. NOX4 mediates hypoxia-induced proliferation of human pulmonary artery smooth muscle cells: the role of autocrine production of transforming growth factor-{beta}1 and insulin-like growth factor binding protein-3. Am J Physiol Lung Cell Mol Physiol. 2009;296(3):L489–99. doi: 10.1152/ajplung.90488.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Djordjevic T, BelAiba RS, Bonello S, Pfeilschifter J, Hess J, Gorlach A. Human urotensin II is a novel activator of NADPH oxidase in human pulmonary artery smooth muscle cells. Arterioscler Thromb Vasc Biol. 2005;25(3):519–25. doi: 10.1161/01.ATV.0000154279.98244.eb. [DOI] [PubMed] [Google Scholar]
  • 100.Schroder K, Wandzioch K, Helmcke I, Brandes RP. Nox4 acts as a switch between differentiation and proliferation in preadipocytes. Arterioscler Thromb Vasc Biol. 2009;29(2):239–45. doi: 10.1161/ATVBAHA.108.174219. [DOI] [PubMed] [Google Scholar]
  • 101.Barnes JL, Gorin Y. Myofibroblast differentiation during fibrosis: role of NAD (P)H oxidases. Kidney Int. 2010;79(9):944–956. doi: 10.1038/ki.2010.516. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Li PF, Dietz R, von Harsdorf R. Differential effect of hydrogen peroxide and superoxide anion on apoptosis and proliferation of vascular smooth muscle cells. Circulation. 1997;96(10):3602–9. doi: 10.1161/01.cir.96.10.3602. [DOI] [PubMed] [Google Scholar]
  • 103.Zafari AM, 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(3):488–95. doi: 10.1161/01.hyp.32.3.488. [DOI] [PubMed] [Google Scholar]
  • 104.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(5234):296–9. doi: 10.1126/science.270.5234.296. [DOI] [PubMed] [Google Scholar]
  • 105.Schroder K. Isoform specific functions of Nox protein-derived reactive oxygen species in the vasculature. Curr Opin Pharmacol. 2010;10(2):122–6. doi: 10.1016/j.coph.2010.01.002. [DOI] [PubMed] [Google Scholar]
  • 106.Brandes RP, Schroder K. Differential vascular functions of Nox family NADPH oxidases. Curr Opin Lipidol. 2008;19(5):513–8. doi: 10.1097/MOL.0b013e32830c91e3. [DOI] [PubMed] [Google Scholar]
  • 107.Irani K. Oxidant signaling in vascular cell growth, death, and survival : a review of the roles of reactive oxygen species in smooth muscle and endothelial cell mitogenic and apoptotic signaling. Circ Res. 2000;87(3):179–83. doi: 10.1161/01.res.87.3.179. [DOI] [PubMed] [Google Scholar]
  • 108.Lee MY, San Martin A, Mehta PK, Dikalova AE, Garrido AM, Datla SR, Lyons E, Krause KH, Banfi B, Lambeth JD, Lassegue B, Griendling KK. Mechanisms of vascular smooth muscle NADPH oxidase 1 (Nox1) contribution to injury-induced neointimal formation. Arterioscler Thromb Vasc Biol. 2009;29(4):480–7. doi: 10.1161/ATVBAHA.108.181925. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Menshikov M, Plekhanova O, Cai H, Chalupsky K, Parfyonova Y, Bashtrikov P, Tkachuk V, Berk BC. Urokinase plasminogen activator stimulates vascular smooth muscle cell proliferation via redox-dependent pathways. Arterioscler Thromb Vasc Biol. 2006;26(4):801–7. doi: 10.1161/01.ATV.0000207277.27432.15. [DOI] [PubMed] [Google Scholar]
  • 110.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(38):23317–21. doi: 10.1074/jbc.271.38.23317. [DOI] [PubMed] [Google Scholar]
  • 111.Ushio-Fukai M, Alexander RW, Akers M, Yin Q, Fujio Y, Walsh K, Griendling KK. Reactive oxygen species mediate the activation of Akt/protein kinase B by angiotensin II in vascular smooth muscle cells. J Biol Chem. 1999;274(32):22699–704. doi: 10.1074/jbc.274.32.22699. [DOI] [PubMed] [Google Scholar]
  • 112.Garrido AM, Griendling KK. NADPH oxidases and angiotensin II receptor signaling. Mol Cell Endocrinol. 2009;302(2):148–58. doi: 10.1016/j.mce.2008.11.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Niu XL, Madamanchi NR, Vendrov AE, Tchivilev I, Rojas M, Madamanchi C, Brandes RP, Krause KH, Humphries J, Smith A, Burnand KG, Runge MS. Nox activator 1: a potential target for modulation of vascular reactive oxygen species in atherosclerotic arteries. Circulation. 2010;121(4):549–59. doi: 10.1161/CIRCULATIONAHA.109.908319. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Jiang B, Xu S, Hou X, Pimentel DR, Cohen RA. Angiotensin II differentially regulates interleukin-1-beta-inducible NO synthase (iNOS) and vascular cell adhesion molecule-1 (VCAM-1) expression: role of p38 MAPK. J Biol Chem. 2004;279(19):20363–8. doi: 10.1074/jbc.M314172200. [DOI] [PubMed] [Google Scholar]
  • 115.Bindoli A, Fukuto JM, Forman HJ. Thiol chemistry in peroxidase catalysis and redox signaling. Antioxid Redox Signal. 2008;10(9):1549–64. doi: 10.1089/ars.2008.2063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Gutscher M, Sobotta MC, Wabnitz GH, Ballikaya S, Meyer AJ, Samstag Y, Dick TP. Proximity-based protein thiol oxidation by H2O2-scavenging peroxidases. J Biol Chem. 2009;284(46):31532–40. doi: 10.1074/jbc.M109.059246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Stanic B, Katsuyama M, Miller FJ., Jr An oxidized extracellular oxidation-reduction state increases Nox1 expression and proliferation in vascular smooth muscle cells via epidermal growth factor receptor activation. Arterioscler Thromb Vasc Biol. 2010;30(11):2234–41. doi: 10.1161/ATVBAHA.110.207639. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Li JM, Shah AM. Intracellular localization and preassembly of the NADPH oxidase complex in cultured endothelial cells. J Biol Chem. 2002;277(22):19952–60. doi: 10.1074/jbc.M110073200. [DOI] [PubMed] [Google Scholar]
  • 119.Van Buul JD, Fernandez-Borja M, Anthony EC, Hordijk PL. Expression and localization of NOX2 and NOX4 in primary human endothelial cells. Antioxid Redox Signal. 2005;7(3–4):308–17. doi: 10.1089/ars.2005.7.308. [DOI] [PubMed] [Google Scholar]
  • 120.Petry A, Djordjevic T, Weitnauer M, Kietzmann T, Hess J, Gorlach A. NOX2 and NOX4 mediate proliferative response in endothelial cells. Antioxid Redox Signal. 2006;8(9–10):1473–84. doi: 10.1089/ars.2006.8.1473. [DOI] [PubMed] [Google Scholar]
  • 121.Datla SR, Peshavariya H, Dusting GJ, Mahadev K, Goldstein BJ, Jiang F. Important role of Nox4 type NADPH oxidase in angiogenic responses in human microvascular endothelial cells in vitro. Arterioscler Thromb Vasc Biol. 2007;27(11):2319–24. doi: 10.1161/ATVBAHA.107.149450. [DOI] [PubMed] [Google Scholar]
  • 122.Brandes RP. Role of NADPH oxidases in the control of vascular gene expression. Antioxid Redox Signal. 2003;5(6):803–11. doi: 10.1089/152308603770380115. [DOI] [PubMed] [Google Scholar]
  • 123.Dourron HM, Jacobson GM, Park JL, Liu J, Reddy DJ, Scheel ML, Pagano PJ. Perivascular gene transfer of NADPH oxidase inhibitor suppresses angioplasty-induced neointimal proliferation of rat carotid artery. Am J Physiol Heart Circ Physiol. 2005;288(2):H946–53. doi: 10.1152/ajpheart.00413.2004. [DOI] [PubMed] [Google Scholar]
  • 124.Haurani MJ, Pagano PJ. Adventitial fibroblast reactive oxygen species as autacrine and paracrine mediators of remodeling: bellwether for vascular disease? Cardiovasc Res. 2007;75(4):679–89. doi: 10.1016/j.cardiores.2007.06.016. [DOI] [PubMed] [Google Scholar]
  • 125.Dickinson BC, Peltier J, Stone D, Schaffer DV, Chang CJ. Nox2 redox signaling maintains essential cell populations in the brain. Nat Chem Biol. 2010;7(2):106–12. doi: 10.1038/nchembio.497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.San Martin A, Griendling KK. Redox control of vascular smooth muscle migration. Antioxid Redox Signal. 2010;12(5):625–40. doi: 10.1089/ars.2009.2852. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Weber DS, Taniyama Y, Rocic P, Seshiah PN, Dechert MA, Gerthoffer WT, Griendling KK. Phosphoinositide-dependent kinase 1 and p21-activated protein kinase mediate reactive oxygen species-dependent regulation of platelet-derived growth factor-induced smooth muscle cell migration. Circ Res. 2004;94(9):1219–26. doi: 10.1161/01.RES.0000126848.54740.4A. [DOI] [PubMed] [Google Scholar]
  • 128.San Martin A, Lee MY, Williams HC, Mizuno K, Lassegue B, Griendling KK. Dual regulation of cofilin activity by LIM kinase and Slingshot-1L phosphatase controls platelet-derived growth factor-induced migration of human aortic smooth muscle cells. Circ Res. 2008;102(4):432–8. doi: 10.1161/CIRCRESAHA.107.158923. [DOI] [PubMed] [Google Scholar]
  • 129.Schroder K, Kohnen A, Aicher A, Liehn EA, Buchse T, Stein S, Weber C, Dimmeler S, Brandes RP. NADPH oxidase Nox2 is required for hypoxia-induced mobilization of endothelial progenitor cells. Circ Res. 2009;105(6):537–44. doi: 10.1161/CIRCRESAHA.109.205138. [DOI] [PubMed] [Google Scholar]
  • 130.Bae YS, Lee JH, Choi SH, Kim S, Almazan F, Witztum JL, Miller YI. Macrophages generate reactive oxygen species in response to minimally oxidized low-density lipoprotein: toll-like receptor 4- and spleen tyrosine kinase-dependent activation of NADPH oxidase 2. Circ Res. 2009;104(2):210–8. doi: 10.1161/CIRCRESAHA.108.181040. 21p following 218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Ushio-Fukai M, Tang Y, Fukai T, Dikalov SI, Ma Y, Fujimoto M, Quinn MT, Pagano PJ, Johnson C, Alexander RW. Novel role of gp91 (phox)-containing NAD (P)H oxidase in vascular endothelial growth factor-induced signaling and angiogenesis. Circ Res. 2002;91(12):1160–7. doi: 10.1161/01.res.0000046227.65158.f8. [DOI] [PubMed] [Google Scholar]
  • 132.Brandes RP, Viedt C, Nguyen K, Beer S, Kreuzer J, Busse R, Gorlach A. Thrombin-induced MCP-1 expression involves activation of the p22phox-containing NADPH oxidase in human vascular smooth muscle cells. Thromb Haemost. 2001;85(6):1104–10. [PubMed] [Google Scholar]
  • 133.ten Freyhaus H, Huntgeburth M, Wingler K, Schnitker J, Baumer AT, Vantler M, Bekhite MM, Wartenberg M, Sauer H, Rosenkranz S. Novel Nox inhibitor VAS2870 attenuates PDGF-dependent smooth muscle cell chemotaxis, but not proliferation. Cardiovasc Res. 2006;71(2):331–41. doi: 10.1016/j.cardiores.2006.01.022. [DOI] [PubMed] [Google Scholar]
  • 134.Rodriguez AI, Gangopadhyay A, Kelley EE, Pagano PJ, Zuckerbraun BS, Bauer PM. HO-1 and CO decrease platelet-derived growth factor-induced vascular smooth muscle cell migration via inhibition of Nox1. Arterioscler Thromb Vasc Biol. 2010;30(1):98–104. doi: 10.1161/ATVBAHA.109.197822. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.San Martin ALMGKK. Novel Nox1-mediated mechanism of SSH1L activation in VSMC. Role in Cell Migration. Atherosclerosis, Thrombosis and Vascular Biology. 2008;28(e109) [Google Scholar]
  • 136.Schroder K, Helmcke I, Palfi K, Krause KH, Busse R, Brandes RP. Nox1 mediates basic fibroblast growth factor-induced migration of vascular smooth muscle cells. Arterioscler Thromb Vasc Biol. 2007;27(8):1736–43. doi: 10.1161/ATVBAHA.107.142117. [DOI] [PubMed] [Google Scholar]
  • 137.Shinohara M, Adachi Y, Mitsushita J, Kuwabara M, Nagasawa A, Harada S, Furuta S, Zhang Y, Seheli K, Miyazaki H, Kamata T. Reactive oxygen generated by NADPH oxidase 1 (Nox1) contributes to cell invasion by regulating matrix metalloprotease-9 production and cell migration. J Biol Chem. 2010;285(7):4481–8. doi: 10.1074/jbc.M109.071779. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Sadok A, Pierres A, Dahan L, Prevot C, Lehmann M, Kovacic H. NADPH oxidase 1 controls the persistence of directed cell migration by a Rho-dependent switch of alpha2/alpha3 integrins. Mol Cell Biol. 2009;29(14):3915–28. doi: 10.1128/MCB.01199-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Meng D, Lv DD, Fang J. Insulin-like growth factor-I induces reactive oxygen species production and cell migration through Nox4 and Rac1 in vascular smooth muscle cells. Cardiovasc Res. 2008;80(2):299–308. doi: 10.1093/cvr/cvn173. [DOI] [PubMed] [Google Scholar]
  • 140.Tobar N, Guerrero J, Smith PC, Martinez J. NOX4-dependent ROS production by stromal mammary cells modulates epithelial MCF-7 cell migration. Br J Cancer. 2010;103(7):1040–7. doi: 10.1038/sj.bjc.6605847. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Haurani MJ, Cifuentes ME, Shepard AD, Pagano PJ. Nox4 oxidase overexpression specifically decreases endogenous Nox4 mRNA and inhibits angiotensin II-induced adventitial myofibroblast migration. Hypertension. 2008;52(1):143–9. doi: 10.1161/HYPERTENSIONAHA.107.101667. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Tong X, Schroder K. NADPH oxidases are responsible for the failure of nitric oxide to inhibit migration of smooth muscle cells exposed to high glucose. Free Radic Biol Med. 2009;47(11):1578–83. doi: 10.1016/j.freeradbiomed.2009.08.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Tong X, Hou X, Jourd’heuil D, Weisbrod RM, Cohen RA. Upregulation of Nox4 by TGF{beta}1 oxidizes SERCA and inhibits NO in arterial smooth muscle of the prediabetic Zucker rat. Circ Res. 2010;107(8):975–83. doi: 10.1161/CIRCRESAHA.110.221242. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Yorimitsu T, Klionsky DJ. Eating the endoplasmic reticulum: quality control by autophagy. Trends Cell Biol. 2007;17(6):279–85. doi: 10.1016/j.tcb.2007.04.005. [DOI] [PubMed] [Google Scholar]
  • 145.Nishida K, Yamaguchi O, Otsu K. Crosstalk between autophagy and apoptosis in heart disease. Circ Res. 2008;103(4):343–51. doi: 10.1161/CIRCRESAHA.108.175448. [DOI] [PubMed] [Google Scholar]
  • 146.Filomeni G, Ciriolo MR. Redox control of apoptosis: an update. Antioxid Redox Signal. 2006;8(11–12):2187–92. doi: 10.1089/ars.2006.8.2187. [DOI] [PubMed] [Google Scholar]
  • 147.Terada LS. Specificity in reactive oxidant signaling: think globally, act locally. J Cell Biol. 2006;174(5):615–23. doi: 10.1083/jcb.200605036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Basuroy S, Bhattacharya S, Leffler CW, Parfenova H. Nox4 NADPH oxidase mediates oxidative stress and apoptosis caused by TNF-alpha in cerebral vascular endothelial cells. Am J Physiol Cell Physiol. 2009;296(3):C422–32. doi: 10.1152/ajpcell.00381.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Dong F, Zhang X, Li SY, Zhang Z, Ren Q, Culver B, Ren J. Possible involvement of NADPH oxidase and JNK in homocysteine-induced oxidative stress and apoptosis in human umbilical vein endothelial cells. Cardiovasc Toxicol. 2005;5(1):9–20. doi: 10.1385/ct:5:1:009. [DOI] [PubMed] [Google Scholar]
  • 150.Shin HK, Kim YK, Kim KY, Lee JH, Hong KW. Remnant lipoprotein particles induce apoptosis in endothelial cells by NAD (P)H oxidase-mediated production of superoxide and cytokines via lectin-like oxidized low-density lipoprotein receptor-1 activation: prevention by cilostazol. Circulation. 2004;109(8):1022–8. doi: 10.1161/01.CIR.0000117403.64398.53. [DOI] [PubMed] [Google Scholar]
  • 151.Van Laethem A, Nys K, Van Kelst S, Claerhout S, Ichijo H, Vandenheede JR, Garmyn M, Agostinis P. Apoptosis signal regulating kinase-1 connects reactive oxygen species to p38 MAPK-induced mitochondrial apoptosis in UVB-irradiated human keratinocytes. Free Radic Biol Med. 2006;41(9):1361–71. doi: 10.1016/j.freeradbiomed.2006.07.007. [DOI] [PubMed] [Google Scholar]
  • 152.Qin F, Patel R, Yan C, Liu W. NADPH oxidase is involved in angiotensin II-induced apoptosis in H9C2 cardiac muscle cells: effects of apocynin. Free Radic Biol Med. 2006;40(2):236–46. doi: 10.1016/j.freeradbiomed.2005.08.010. [DOI] [PubMed] [Google Scholar]
  • 153.Ago T, Kuroda J, Pain J, Fu C, Li H, Sadoshima J. Upregulation of Nox4 by hypertrophic stimuli promotes apoptosis and mitochondrial dysfunction in cardiac myocytes. Circ Res. 2010;106(7):1253–64. doi: 10.1161/CIRCRESAHA.109.213116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Kuroda J, Ago T, Matsushima S, Zhai P, Schneider MD, Sadoshima J. NADPH oxidase 4 (Nox4) is a major source of oxidative stress in the failing heart. Proc Natl Acad Sci U S A. 2010;107(35):15565–70. doi: 10.1073/pnas.1002178107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Johnson TM, Yu ZX, Ferrans VJ, Lowenstein RA, Finkel T. Reactive oxygen species are downstream mediators of p53-dependent apoptosis. Proc Natl Acad Sci U S A. 1996;93(21):11848–52. doi: 10.1073/pnas.93.21.11848. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Speir E, Modali R, Huang ES, Leon MB, Shawl F, Finkel T, Epstein SE. Potential role of human cytomegalovirus and p53 interaction in coronary restenosis. Science. 1994;265(5170):391–4. doi: 10.1126/science.8023160. [DOI] [PubMed] [Google Scholar]
  • 157.Li PF, Dietz R, von Harsdorf R. Reactive oxygen species induce apoptosis of vascular smooth muscle cell. FEBS Lett. 1997;404(2–3):249–52. doi: 10.1016/s0014-5793(97)00093-8. [DOI] [PubMed] [Google Scholar]
  • 158.Li Y, Zhang F, Nagai N, Tang Z, Zhang S, Scotney P, Lennartsson J, Zhu C, Qu Y, Fang C, Hua J, Matsuo O, Fong GH, Ding H, Cao Y, Becker KG, Nash A, Heldin CH, Li X. VEGF-B inhibits apoptosis via VEGFR-1-mediated suppression of the expression of BH3-only protein genes in mice and rats. J Clin Invest. 2008;118(3):913–23. doi: 10.1172/JCI33673. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Rutkowski DT, Wu J, Back SH, Callaghan MU, Ferris SP, Iqbal J, Clark R, Miao H, Hassler JR, Fornek J, Katze MG, Hussain MM, Song B, Swathirajan J, Wang J, Yau GD, Kaufman RJ. UPR pathways combine to prevent hepatic steatosis caused by ER stress-mediated suppression of transcriptional master regulators. Dev Cell. 2008;15(6):829–40. doi: 10.1016/j.devcel.2008.10.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Rutkowski DT, Arnold SM, Miller CN, Wu J, Li J, Gunnison KM, Mori K, Sadighi Akha AA, Raden D, Kaufman RJ. Adaptation to ER stress is mediated by differential stabilities of pro-survival and pro-apoptotic mRNAs and proteins. PLoS Biol. 2006;4(11):e374. doi: 10.1371/journal.pbio.0040374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Minamino T, Kitakaze M. ER stress in cardiovascular disease. J Mol Cell Cardiol. 2010;48(6):1105–10. doi: 10.1016/j.yjmcc.2009.10.026. [DOI] [PubMed] [Google Scholar]
  • 162.Kitakaze M, Tsukamoto O. What is the role of ER stress in the heart? Introduction and series overview. Circ Res. 2010;107(1):15–8. doi: 10.1161/CIRCRESAHA.110.222919. [DOI] [PubMed] [Google Scholar]
  • 163.Heath-Engel HM, Chang NC, Shore GC. The endoplasmic reticulum in apoptosis and autophagy: role of the BCL-2 protein family. Oncogene. 2008;27(50):6419–33. doi: 10.1038/onc.2008.309. [DOI] [PubMed] [Google Scholar]
  • 164.Merksamer PI, Trusina A, Papa FR. Real-time redox measurements during endoplasmic reticulum stress reveal interlinked protein folding functions. Cell. 2008;135(5):933–47. doi: 10.1016/j.cell.2008.10.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Malhotra JD, Kaufman RJ. Endoplasmic reticulum stress and oxidative stress: a vicious cycle or a double-edged sword? Antioxid Redox Signal. 2007;9(12):2277–93. doi: 10.1089/ars.2007.1782. [DOI] [PubMed] [Google Scholar]
  • 166.Amanso AM, Debbas V, Laurindo FR. Proteasome inhibition represses unfolded protein response and Nox4, sensitizing vascular cells to endoplasmic reticulum stress-induced death. PLoS One. 2011;6(1):e14591. doi: 10.1371/journal.pone.0014591. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Younce CW, Kolattukudy PE. MCP-1 causes cardiomyoblast death via autophagy resulting from ER stress caused by oxidative stress generated by inducing a novel zinc-finger protein, MCPIP. Biochem J. 2010;426(1):43–53. doi: 10.1042/BJ20090976. [DOI] [PubMed] [Google Scholar]
  • 168.Li G, Scull C, Ozcan L, Tabas I. NADPH oxidase links endoplasmic reticulum stress, oxidative stress, and PKR activation to induce apoptosis. J Cell Biol. 2010;191(6):1113–25. doi: 10.1083/jcb.201006121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Loughlin DT, Artlett CM. Precursor of advanced glycation end products mediates ER-stress-induced caspase-3 activation of human dermal fibroblasts through NAD (P)H oxidase 4. PLoS One. 2011;5(6):e11093. doi: 10.1371/journal.pone.0011093. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Li J, Zhu H, Shen E, Wan L, Arnold JM, Peng T. Deficiency of rac1 blocks NADPH oxidase activation, inhibits endoplasmic reticulum stress, and reduces myocardial remodeling in a mouse model of type 1 diabetes. Diabetes. 2010;59(8):2033–42. doi: 10.2337/db09-1800. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Wu RF, Ma Z, Liu Z, Terada LS. Nox4-derived H2O2 mediates endoplasmic reticulum signaling through local Ras activation. Mol Cell Biol. 2010;30(14):3553–68. doi: 10.1128/MCB.01445-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Pedruzzi E, Guichard C, Ollivier V, Driss F, Fay M, Prunet C, Marie JC, Pouzet C, Samadi M, Elbim C, O’Dowd Y, Bens M, Vandewalle A, Gougerot-Pocidalo MA, Lizard G, Ogier-Denis E. NAD (P)H oxidase Nox-4 mediates 7-ketocholesterol-induced endoplasmic reticulum stress and apoptosis in human aortic smooth muscle cells. Mol Cell Biol. 2004;24(24):10703–17. doi: 10.1128/MCB.24.24.10703-10717.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Tabas I, Ron D. Integrating the mechanisms of apoptosis induced by endoplasmic reticulum stress. Nat Cell Biol. 2011;13(3):184–90. doi: 10.1038/ncb0311-184. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Gu F, Nguyen DT, Stuible M, Dube N, Tremblay ML, Chevet E. Protein-tyrosine phosphatase 1B potentiates IRE1 signaling during endoplasmic reticulum stress. J Biol Chem. 2004;279(48):49689–93. doi: 10.1074/jbc.C400261200. [DOI] [PubMed] [Google Scholar]
  • 175.Chen K, Kirber MT, Xiao H, Yang Y, Keaney JF., Jr Regulation of ROS signal transduction by NADPH oxidase 4 localization. J Cell Biol. 2008;181(7):1129–39. doi: 10.1083/jcb.200709049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Ben Mkaddem S, Pedruzzi E, Werts C, Coant N, Bens M, Cluzeaud F, Goujon JM, Ogier-Denis E, Vandewalle A. Heat shock protein gp96 and NAD (P)H oxidase 4 play key roles in Toll-like receptor 4-activated apoptosis during renal ischemia/reperfusion injury. Cell Death Differ. 2010;17(9):1474–85. doi: 10.1038/cdd.2010.26. [DOI] [PubMed] [Google Scholar]
  • 177.Huang J, Brumell JH. NADPH oxidases contribute to autophagy regulation. Autophagy. 2009;5(6):887–9. doi: 10.4161/auto.9125. [DOI] [PubMed] [Google Scholar]
  • 178.Li J, Ni M, Lee B, Barron E, Hinton DR, Lee AS. The unfolded protein response regulator GRP78/BiP is required for endoplasmic reticulum integrity and stress-induced autophagy in mammalian cells. Cell Death Differ. 2008;15(9):1460–71. doi: 10.1038/cdd.2008.81. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Huang J, Canadien V, Lam GY, Steinberg BE, Dinauer MC, Magalhaes MA, Glogauer M, Grinstein S, Brumell JH. Activation of antibacterial autophagy by NADPH oxidases. Proc Natl Acad Sci U S A. 2009;106(15):6226–31. doi: 10.1073/pnas.0811045106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Semenza GL. Life with oxygen. Science. 2007;318(5847):62–4. doi: 10.1126/science.1147949. [DOI] [PubMed] [Google Scholar]
  • 181.Semenza GL. Regulation of oxygen homeostasis by hypoxia-inducible factor 1. Physiology (Bethesda) 2009;24:97–106. doi: 10.1152/physiol.00045.2008. [DOI] [PubMed] [Google Scholar]
  • 182.Majmundar AJ, Wong WJ, Simon MC. Hypoxia-inducible factors and the response to hypoxic stress. Mol Cell. 2010;40(2):294–309. doi: 10.1016/j.molcel.2010.09.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Maxwell PH, Wiesener MS, Chang GW, Clifford SC, Vaux EC, Cockman ME, Wykoff CC, Pugh CW, Maher ER, Ratcliffe PJ. The tumour suppressor protein VHL targets hypoxia-inducible factors for oxygen-dependent proteolysis. Nature. 1999;399(6733):271–5. doi: 10.1038/20459. [DOI] [PubMed] [Google Scholar]
  • 184.Kelly BD, Hackett SF, Hirota K, Oshima Y, Cai Z, Berg-Dixon S, Rowan A, Yan Z, Campochiaro PA, Semenza GL. Cell type-specific regulation of angiogenic growth factor gene expression and induction of angiogenesis in nonischemic tissue by a constitutively active form of hypoxia-inducible factor 1. Circ Res. 2003;93(11):1074–81. doi: 10.1161/01.RES.0000102937.50486.1B. [DOI] [PubMed] [Google Scholar]
  • 185.Bosch-Marce M, Okuyama H, Wesley JB, Sarkar K, Kimura H, Liu YV, Zhang H, Strazza M, Rey S, Savino L, Zhou YF, McDonald KR, Na Y, Vandiver S, Rabi A, Shaked Y, Kerbel R, Lavallee T, Semenza GL. Effects of aging and hypoxia-inducible factor-1 activity on angiogenic cell mobilization and recovery of perfusion after limb ischemia. Circ Res. 2007;101(12):1310–8. doi: 10.1161/CIRCRESAHA.107.153346. [DOI] [PubMed] [Google Scholar]
  • 186.Fu XW, Wang D, Nurse CA, Dinauer MC, Cutz E. NADPH oxidase is an O2 sensor in airway chemoreceptors: evidence from K+ current modulation in wild-type and oxidase-deficient mice. Proc Natl Acad Sci U S A. 2000;97(8):4374–9. doi: 10.1073/pnas.97.8.4374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Archer SL, Reeve HL, Michelakis E, Puttagunta L, Waite R, Nelson DP, Dinauer MC, Weir EK. O2 sensing is preserved in mice lacking the gp91 phox subunit of NADPH oxidase. Proc Natl Acad Sci U S A. 1999;96(14):7944–9. doi: 10.1073/pnas.96.14.7944. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Goyal P, Weissmann N, Grimminger F, Hegel C, Bader L, Rose F, Fink L, Ghofrani HA, Schermuly RT, Schmidt HH, Seeger W, Hanze J. Upregulation of NAD (P)H oxidase 1 in hypoxia activates hypoxia-inducible factor 1 via increase in reactive oxygen species. Free Radic Biol Med. 2004;36(10):1279–88. doi: 10.1016/j.freeradbiomed.2004.02.071. [DOI] [PubMed] [Google Scholar]
  • 189.Diebold I, Petry A, Hess J, Gorlach A. The NADPH oxidase subunit NOX4 is a new target gene of the hypoxia-inducible factor-1. Mol Biol Cell. 2010;21(12):2087–96. doi: 10.1091/mbc.E09-12-1003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Park SJ, Chun YS, Park KS, Kim SJ, Choi SO, Kim HL, Park JW. Identification of subdomains in NADPH oxidase-4 critical for the oxygen-dependent regulation of TASK-1 K+ channels. Am J Physiol Cell Physiol. 2009;297(4):C855–64. doi: 10.1152/ajpcell.00463.2008. [DOI] [PubMed] [Google Scholar]
  • 191.Lee YM, Kim BJ, Chun YS, So I, Choi H, Kim MS, Park JW. NOX4 as an oxygen sensor to regulate TASK-1 activity. Cell Signal. 2006;18(4):499–507. doi: 10.1016/j.cellsig.2005.05.025. [DOI] [PubMed] [Google Scholar]
  • 192.Stenmark KR, Fagan KA, Frid MG. Hypoxia-induced pulmonary vascular remodeling: cellular and molecular mechanisms. Circ Res. 2006;99(7):675–91. doi: 10.1161/01.RES.0000243584.45145.3f. [DOI] [PubMed] [Google Scholar]
  • 193.Coggins MP, Bloch KD. Nitric oxide in the pulmonary vasculature. Arterioscler Thromb Vasc Biol. 2007;27(9):1877–85. doi: 10.1161/ATVBAHA.107.142943. [DOI] [PubMed] [Google Scholar]
  • 194.Liu JQ, Sham JS, Shimoda LA, Kuppusamy P, Sylvester JT. Hypoxic constriction and reactive oxygen species in porcine distal pulmonary arteries. Am J Physiol Lung Cell Mol Physiol. 2003;285(2):L322–33. doi: 10.1152/ajplung.00337.2002. [DOI] [PubMed] [Google Scholar]
  • 195.Van Rheen Z, Fattman C, Domarski S, Majka S, Klemm D, Stenmark KR, Nozik-Grayck E. Lung extracellular superoxide dismutase overexpression lessens bleomycin-induced pulmonary hypertension and vascular remodeling. Am J Respir Cell Mol Biol. 2011;44(4):500–8. doi: 10.1165/rcmb.2010-0065OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Dennis KE, Aschner JL, Milatovic D, Schmidt JW, Aschner M, Kaplowitz MR, Zhang Y, Fike CD. NADPH oxidases and reactive oxygen species at different stages of chronic hypoxia-induced pulmonary hypertension in newborn piglets. Am J Physiol Lung Cell Mol Physiol. 2009;297(4):L596–607. doi: 10.1152/ajplung.90568.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Liu JQ, Erbynn EM, Folz RJ. Chronic hypoxia-enhanced murine pulmonary vasoconstriction: role of superoxide and gp91phox. Chest. 2005;128(6 Suppl):594S–596S. doi: 10.1378/chest.128.6_suppl.594S. [DOI] [PubMed] [Google Scholar]
  • 198.Sanders KA, Sundar KM, He L, Dinger B, Fidone S, Hoidal JR. Role of components of the phagocytic NADPH oxidase in oxygen sensing. J Appl Physiol. 2002;93(4):1357–64. doi: 10.1152/japplphysiol.00564.2001. [DOI] [PubMed] [Google Scholar]
  • 199.Fresquet F, Pourageaud F, Leblais V, Brandes RP, Savineau JP, Marthan R, Muller B. Role of reactive oxygen species and gp91phox in endothelial dysfunction of pulmonary arteries induced by chronic hypoxia. Br J Pharmacol. 2006;148(5):714–23. doi: 10.1038/sj.bjp.0706779. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Nisbet RE, Graves AS, Kleinhenz DJ, Rupnow HL, Reed AL, Fan TH, Mitchell PO, Sutliff RL, Hart CM. The role of NADPH oxidase in chronic intermittent hypoxia-induced pulmonary hypertension in mice. Am J Respir Cell Mol Biol. 2009;40(5):601–9. doi: 10.1165/rcmb.2008-0145OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Mittal M, Roth M, Konig P, Hofmann S, Dony E, Goyal P, Selbitz AC, Schermuly RT, Ghofrani HA, Kwapiszewska G, Kummer W, Klepetko W, Hoda MA, Fink L, Hanze J, Seeger W, Grimminger F, Schmidt HH, Weissmann N. Hypoxia-dependent regulation of nonphagocytic NADPH oxidase subunit NOX4 in the pulmonary vasculature. Circ Res. 2007;101(3):258–67. doi: 10.1161/CIRCRESAHA.107.148015. [DOI] [PubMed] [Google Scholar]
  • 202.Ahmad M, Kelly MR, Zhao X, Kandhi S, Wolin MS. Roles for Nox4 in the contractile response of bovine pulmonary arteries to hypoxia. Am J Physiol Heart Circ Physiol. 298(6):H1879–88. doi: 10.1152/ajpheart.01228.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Diebold I, Petry A, Hess J, Gorlach A. The NADPH oxidase subunit NOX4 is a new target gene of the hypoxia-inducible factor-1. Mol Biol Cell. 2010;21(12):2087–96. doi: 10.1091/mbc.E09-12-1003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Diebold I, Flugel D, Becht S, Belaiba RS, Bonello S, Hess J, Kietzmann T, Gorlach A. The hypoxia-inducible factor-2alpha is stabilized by oxidative stress involving NOX4. Antioxid Redox Signal. 2010;13(4):425–36. doi: 10.1089/ars.2009.3014. [DOI] [PubMed] [Google Scholar]
  • 205.Lu X, Murphy TC, Nanes MS, Hart CM. PPAR{gamma} regulates hypoxia-induced Nox4 expression in human pulmonary artery smooth muscle cells through NF-{kappa}B. Am J Physiol Lung Cell Mol Physiol. 2010;299(4):L559–66. doi: 10.1152/ajplung.00090.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Lusis AJ. Atherosclerosis. Nature. 2000;407(6801):233–41. doi: 10.1038/35025203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Heinloth A, Heermeier K, Raff U, Wanner C, Galle J. Stimulation of NADPH oxidase by oxidized low-density lipoprotein induces proliferation of human vascular endothelial cells. J Am Soc Nephrol. 2000;11(10):1819–25. doi: 10.1681/ASN.V11101819. [DOI] [PubMed] [Google Scholar]
  • 208.Aviram M, Rosenblat M, Etzioni A, Levy R. Activation of NADPH oxidase required for macrophage-mediated oxidation of low-density lipoprotein. Metabolism. 1996;45(9):1069–79. doi: 10.1016/s0026-0495(96)90005-0. [DOI] [PubMed] [Google Scholar]
  • 209.Barry-Lane PA, Patterson C, van der Merwe M, Hu Z, Holland SM, Yeh ET, Runge MS. p47phox is required for atherosclerotic lesion progression in ApoE (−/−) mice. J Clin Invest. 2001;108(10):1513–22. doi: 10.1172/JCI11927. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Hsich E, Segal BH, Pagano PJ, Rey FE, Paigen B, Deleonardis J, Hoyt RF, Holland SM, Finkel T. Vascular effects following homozygous disruption of p47 (phox) : An essential component of NADPH oxidase. Circulation. 2000;101(11):1234–6. doi: 10.1161/01.cir.101.11.1234. [DOI] [PubMed] [Google Scholar]
  • 211.Sheehan AL, Carrell S, Johnson B, Stanic B, Banfi B, Miller FJ., Jr Role for Nox1 NADPH oxidase in atherosclerosis. Atherosclerosis. 2011 doi: 10.1016/j.atherosclerosis.2011.02.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Zhou MS, Hernandez Schulman I, Pagano PJ, Jaimes EA, Raij L. Reduced NAD (P)H oxidase in low renin hypertension: link among angiotensin II, atherogenesis, and blood pressure. Hypertension. 2006;47(1):81–6. doi: 10.1161/01.HYP.0000197182.65554.c7. [DOI] [PubMed] [Google Scholar]
  • 213.Tabas I. The Role of Endoplasmic Reticulum Stress in the Progression of Atherosclerosis. Circulation Research. 2010;107(7):839–850. doi: 10.1161/CIRCRESAHA.110.224766. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214.De Meyer GRY, Martinet W. Autophagy in the cardiovascular system. Biochimica Et Biophysica Acta-Molecular Cell Research. 2009;1793(9):1485–1495. doi: 10.1016/j.bbamcr.2008.12.011. [DOI] [PubMed] [Google Scholar]
  • 215.Martinet W, De Meyer GRY. Autophagy in Atherosclerosis A Cell Survival and Death Phenomenon With Therapeutic Potential. Circulation Research. 2009;104(3):304–317. doi: 10.1161/CIRCRESAHA.108.188318. [DOI] [PubMed] [Google Scholar]
  • 216.Myoishi M, Hao H, Minamino T, Watanabe K, Nishihira K, Hatakeyama K, Asada Y, Okada KI, Ishibashi-Ueda H, Gabbiani G, Bochaton-Piallat ML, Mochizuki N, Kitakaze M. Increased endoplasmic reticulum stress in atherosclerotic plaques associated with acute coronary syndrome. Circulation. 2007;116(11):1226–1233. doi: 10.1161/CIRCULATIONAHA.106.682054. [DOI] [PubMed] [Google Scholar]
  • 217.Thorp E, Li G, Seimon TA, Kuriakose G, Ron D, Tabas I. Reduced apoptosis and plaque necrosis in advanced atherosclerotic lesions of Apoe−/− and Ldlr−/− mice lacking CHOP. Cell Metab. 2009;9(5):474–81. doi: 10.1016/j.cmet.2009.03.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Ron D, Walter P. Signal integration in the endoplasmic reticulum unfolded protein response. Nat Rev Mol Cell Biol. 2007;8(7):519–29. doi: 10.1038/nrm2199. [DOI] [PubMed] [Google Scholar]
  • 219.Park HS, Jung HY, Park EY, Kim J, Lee WJ, Bae YS. Cutting edge: direct interaction of TLR4 with NAD (P)H oxidase 4 isozyme is essential for lipopolysaccharide-induced production of reactive oxygen species and activation of NF-kappa B. J Immunol. 2004;173(6):3589–93. doi: 10.4049/jimmunol.173.6.3589. [DOI] [PubMed] [Google Scholar]
  • 220.Park HS, Chun JN, Jung HY, Choi C, Bae YS. Role of NADPH oxidase 4 in lipopolysaccharide-induced proinflammatory responses by human aortic endothelial cells. Cardiovasc Res. 2006;72(3):447–55. doi: 10.1016/j.cardiores.2006.09.012. [DOI] [PubMed] [Google Scholar]
  • 221.Maloney E, Sweet IR, Hockenbery DM, Pham M, Rizzo NO, Tateya S, Handa P, Schwartz MW, Kim F. Activation of NF-kappaB by palmitate in endothelial cells: a key role for NADPH oxidase-derived superoxide in response to TLR4 activation. Arterioscler Thromb Vasc Biol. 2009;29(9):1370–5. doi: 10.1161/ATVBAHA.109.188813. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Cai H, Griendling KK, Harrison DG. The vascular NAD (P)H oxidases as therapeutic targets in cardiovascular diseases. Trends Pharmacol Sci. 2003;24(9):471–8. doi: 10.1016/S0165-6147(03)00233-5. [DOI] [PubMed] [Google Scholar]
  • 223.Khatri JJ, Johnson C, Magid R, Lessner SM, Laude KM, Dikalov SI, Harrison DG, Sung HJ, Rong Y, Galis ZS. Vascular oxidant stress enhances progression and angiogenesis of experimental atheroma. Circulation. 2004;109(4):520–5. doi: 10.1161/01.CIR.0000109698.70638.2B. [DOI] [PubMed] [Google Scholar]
  • 224.Szocs K, Lassegue 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(1):21–7. doi: 10.1161/hq0102.102189. [DOI] [PubMed] [Google Scholar]
  • 225.Diebold I, Djordjevic T, Petry A, Hatzelmann A, Tenor H, Hess J, Gorlach A. Phosphodiesterase 2 mediates redox-sensitive endothelial cell proliferation and angiogenesis by thrombin via Rac1 and NADPH oxidase 2. Circ Res. 2009;104(10):1169–77. doi: 10.1161/CIRCRESAHA.109.196592. [DOI] [PubMed] [Google Scholar]
  • 226.Al-Shabrawey M, Bartoli M, El-Remessy AB, Platt DH, Matragoon S, Behzadian MA, Caldwell RW, Caldwell RB. Inhibition of NAD (P)H oxidase activity blocks vascular endothelial growth factor overexpression and neovascularization during ischemic retinopathy. Am J Pathol. 2005;167(2):599–607. doi: 10.1016/S0002-9440(10)63001-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Tojo T, Ushio-Fukai M, Yamaoka-Tojo M, Ikeda S, Patrushev N, Alexander RW. Role of gp91phox (Nox2)-containing NAD (P)H oxidase in angiogenesis in response to hindlimb ischemia. Circulation. 2005;111(18):2347–55. doi: 10.1161/01.CIR.0000164261.62586.14. [DOI] [PubMed] [Google Scholar]
  • 228.Urao N, Inomata H, Razvi M, Kim HW, Wary K, McKinney R, Fukai T, Ushio-Fukai M. Role of nox2-based NADPH oxidase in bone marrow and progenitor cell function involved in neovascularization induced by hindlimb ischemia. Circ Res. 2008;103(2):212–20. doi: 10.1161/CIRCRESAHA.108.176230. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Straub AC, Clark KA, Ross MA, Chandra AG, Li S, Gao X, Pagano PJ, Stolz DB, Barchowsky A. Arsenic-stimulated liver sinusoidal capillarization in mice requires NADPH oxidase-generated superoxide. J Clin Invest. 2008;118(12):3980–9. doi: 10.1172/JCI35092. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Arbiser JL, Petros J, Klafter R, Govindajaran B, McLaughlin ER, Brown LF, Cohen C, Moses M, Kilroy S, Arnold RS, Lambeth JD. Reactive oxygen generated by Nox1 triggers the angiogenic switch. Proc Natl Acad Sci U S A. 2002;99(2):715–20. doi: 10.1073/pnas.022630199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 231.Garrido-Urbani S, Jemelin S, Deffert C, Carnesecchi S, Basset O, Szyndralewiez C, Heitz F, Page P, Montet X, Michalik L, Arbiser J, Ruegg C, Krause KH, Imhof B. Targeting Vascular NADPH Oxidase 1 Blocks Tumor Angiogenesis through a PPARalpha Mediated Mechanism. PLoS One. 2011;6(2):e14665. doi: 10.1371/journal.pone.0014665. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232.Reed R, Kolz C, Potter B, Rocic P. The mechanistic basis for the disparate effects of angiotensin II on coronary collateral growth. Arterioscler Thromb Vasc Biol. 2008;28(1):61–7. doi: 10.1161/ATVBAHA.107.154294. [DOI] [PubMed] [Google Scholar]
  • 233.Teng RJ, Eis A, Bakhutashvili I, Arul N, Konduri GG. Increased superoxide production contributes to the impaired angiogenesis of fetal pulmonary arteries with in utero pulmonary hypertension. Am J Physiol Lung Cell Mol Physiol. 2009;297(1):L184–95. doi: 10.1152/ajplung.90455.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234.Sonta T, Inoguchi T, Tsubouchi H, Sekiguchi N, Kobayashi K, Matsumoto S, Utsumi H, Nawata H. Evidence for contribution of vascular NAD (P)H oxidase to increased oxidative stress in animal models of diabetes and obesity. Free Radic Biol Med. 2004;37(1):115–23. doi: 10.1016/j.freeradbiomed.2004.04.001. [DOI] [PubMed] [Google Scholar]
  • 235.Furukawa S, Fujita T, Shimabukuro M, Iwaki M, Yamada Y, Nakajima Y, Nakayama O, Makishima M, Matsuda M, Shimomura I. Increased oxidative stress in obesity and its impact on metabolic syndrome. J Clin Invest. 2004;114(12):1752–61. doi: 10.1172/JCI21625. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 236.San Martin A, Du P, Dikalova A, Lassegue B, Aleman M, Gongora MC, Brown K, Joseph G, Harrison DG, Taylor WR, Jo H, Griendling KK. Reactive oxygen species-selective regulation of aortic inflammatory gene expression in Type 2 diabetes. Am J Physiol Heart Circ Physiol. 2007;292(5):H2073–82. doi: 10.1152/ajpheart.00943.2006. [DOI] [PubMed] [Google Scholar]
  • 237.San Martin A, Foncea R, Laurindo FR, Ebensperger R, Griendling KK, Leighton F. Nox1-based NADPH oxidase-derived superoxide is required for VSMC activation by advanced glycation end-products. Free Radic Biol Med. 2007;42(11):1671–9. doi: 10.1016/j.freeradbiomed.2007.02.002. [DOI] [PubMed] [Google Scholar]
  • 238.Privratsky JR, Wold LE, Sowers JR, Quinn MT, Ren J. AT1 blockade prevents glucose-induced cardiac dysfunction in ventricular myocytes: role of the AT1 receptor and NADPH oxidase. Hypertension. 2003;42(2):206–12. doi: 10.1161/01.HYP.0000082814.62655.85. [DOI] [PubMed] [Google Scholar]
  • 239.Zhang L, Zalewski A, Liu Y, Mazurek T, Cowan S, Martin JL, Hofmann SM, Vlassara H, Shi Y. Diabetes-induced oxidative stress and low-grade inflammation in porcine coronary arteries. Circulation. 2003;108(4):472–8. doi: 10.1161/01.CIR.0000080378.96063.23. [DOI] [PubMed] [Google Scholar]
  • 240.Liu S, Ma X, Gong M, Shi L, Lincoln T, Wang S. Glucose down-regulation of cGMP-dependent protein kinase I expression in vascular smooth muscle cells involves NAD (P)H oxidase-derived reactive oxygen species. Free Radic Biol Med. 2007;42(6):852–63. doi: 10.1016/j.freeradbiomed.2006.12.025. [DOI] [PubMed] [Google Scholar]
  • 241.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(6779):787–90. doi: 10.1038/35008121. [DOI] [PubMed] [Google Scholar]
  • 242.Hink U, Li H, Mollnau H, Oelze M, Matheis E, Hartmann M, Skatchkov M, Thaiss F, Stahl RA, Warnholtz A, Meinertz T, Griendling K, Harrison DG, Forstermann U, Munzel T. Mechanisms underlying endothelial dysfunction in diabetes mellitus. Circ Res. 2001;88(2):E14–22. doi: 10.1161/01.res.88.2.e14. [DOI] [PubMed] [Google Scholar]
  • 243.Ohshiro Y, Ma RC, Yasuda Y, Hiraoka-Yamamoto J, Clermont AC, Isshiki K, Yagi K, Arikawa E, Kern TS, King GL. Reduction of diabetes-induced oxidative stress, fibrotic cytokine expression, and renal dysfunction in protein kinase Cbeta-null mice. Diabetes. 2006;55(11):3112–20. doi: 10.2337/db06-0895. [DOI] [PubMed] [Google Scholar]
  • 244.Taylor PD, McConnell J, Khan IY, Holemans K, Lawrence KM, Asare-Anane H, Persaud SJ, Jones PM, Petrie L, Hanson MA, Poston L. Impaired glucose homeostasis and mitochondrial abnormalities in offspring of rats fed a fat-rich diet in pregnancy. Am J Physiol Regul Integr Comp Physiol. 2005;288(1):R134–9. doi: 10.1152/ajpregu.00355.2004. [DOI] [PubMed] [Google Scholar]
  • 245.Kwan J, Wang H, Munk S, Xia L, Goldberg HJ, Whiteside CI. In high glucose protein kinase C-zeta activation is required for mesangial cell generation of reactive oxygen species. Kidney Int. 2005;68(6):2526–41. doi: 10.1111/j.1523-1755.2005.00660.x. [DOI] [PubMed] [Google Scholar]
  • 246.Yuan H, Lu Y, Huang X, He Q, Man Y, Zhou Y, Wang S, Li J. Suppression of NADPH oxidase 2 substantially restores glucose-induced dysfunction of pancreatic NIT-1 cells. FEBS J. 2010;277(24):5061–71. doi: 10.1111/j.1742-4658.2010.07911.x. [DOI] [PubMed] [Google Scholar]
  • 247.Aguiari P, Leo S, Zavan B, Vindigni V, Rimessi A, Bianchi K, Franzin C, Cortivo R, Rossato M, Vettor R, Abatangelo G, Pozzan T, Pinton P, Rizzuto R. High glucose induces adipogenic differentiation of muscle-derived stem cells. Proc Natl Acad Sci U S A. 2008;105(4):1226–31. doi: 10.1073/pnas.0711402105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 248.Giacco F, Brownlee M. Oxidative stress and diabetic complications. Circ Res. 2010;107(9):1058–70. doi: 10.1161/CIRCRESAHA.110.223545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249.Subasinghe W, Syed I, Kowluru A. Phagocyte-like NADPH oxidase promotes cytokine-induced mitochondrial dysfunction in pancreatic beta-cells: evidence for regulation by Rac1. Am J Physiol Regul Integr Comp Physiol. 2011;300(1):R12–20. doi: 10.1152/ajpregu.00421.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 250.San Jose G, Bidegain J, Robador PA, Diez J, Fortuno A, Zalba G. Insulin-induced NADPH oxidase activation promotes proliferation and matrix metalloproteinase activation in monocytes/macrophages. Free Radic Biol Med. 2009;46(8):1058–67. doi: 10.1016/j.freeradbiomed.2009.01.009. [DOI] [PubMed] [Google Scholar]
  • 251.Goldstein BJ, Mahadev K, Wu X, Zhu L, Motoshima H. Role of insulin-induced reactive oxygen species in the insulin signaling pathway. Antioxid Redox Signal. 2005;7(7–8):1021–31. doi: 10.1089/ars.2005.7.1021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 252.Tong X, Hou X, Jourd’heuil D, Weisbrod RM, Cohen RA. Upregulation of Nox4 by TGF{beta}1 oxidizes SERCA and inhibits NO in arterial smooth muscle of the prediabetic Zucker rat. Circ Res. 2010;107(8):975–83. doi: 10.1161/CIRCRESAHA.110.221242. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253.Ihm SH, Lee JO, Kim SJ, Seung KB, Schini-Kerth VB, Chang K, Oak MH. Catechin prevents endothelial dysfunction in the prediabetic stage of OLETF rats by reducing vascular NADPH oxidase activity and expression. Atherosclerosis. 2009;206(1):47–53. doi: 10.1016/j.atherosclerosis.2009.01.036. [DOI] [PubMed] [Google Scholar]
  • 254.Wautier MP, Chappey O, Corda S, Stern DM, Schmidt AM, Wautier JL. Activation of NADPH oxidase by AGE links oxidant stress to altered gene expression via RAGE. Am J Physiol Endocrinol Metab. 2001;280(5):E685–94. doi: 10.1152/ajpendo.2001.280.5.E685. [DOI] [PubMed] [Google Scholar]
  • 255.Anderson MM, Heinecke JW. Production of N (epsilon)- (carboxymethyl)lysine is impaired in mice deficient in NADPH oxidase: a role for phagocyte-derived oxidants in the formation of advanced glycation end products during inflammation. Diabetes. 2003;52(8):2137–43. doi: 10.2337/diabetes.52.8.2137. [DOI] [PubMed] [Google Scholar]
  • 256.Zhang H, Morgan B, Potter BJ, Ma L, Dellsperger KC, Ungvari Z, Zhang C. Resveratrol improves left ventricular diastolic relaxation in type 2 diabetes by inhibiting oxidative/nitrative stress: in vivo demonstration with magnetic resonance imaging. Am J Physiol Heart Circ Physiol. 2010;299(4):H985–94. doi: 10.1152/ajpheart.00489.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES