Next Article in Journal
Eucalyptol Inhibits Amyloid-β-Induced Barrier Dysfunction in Glucose-Exposed Retinal Pigment Epithelial Cells and Diabetic Eyes
Next Article in Special Issue
Reactive Oxygen Species and Oxidative Stress in the Pathogenesis and Progression of Genetic Diseases of the Connective Tissue
Previous Article in Journal
Solid Lipid Nanoparticles as Carriers of Natural Phenolic Compounds
Previous Article in Special Issue
Regulation of Vascular Calcification by Reactive Oxygen Species
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Vasoconstrictor Mechanisms in Chronic Hypoxia-Induced Pulmonary Hypertension: Role of Oxidant Signaling

Vascular Physiology Group, Department of Cell Biology and Physiology, University of New Mexico Health Sciences Center, Albuquerque, NM 87131, USA
*
Author to whom correspondence should be addressed.
Antioxidants 2020, 9(10), 999; https://doi.org/10.3390/antiox9100999
Submission received: 5 September 2020 / Revised: 6 October 2020 / Accepted: 6 October 2020 / Published: 15 October 2020
(This article belongs to the Special Issue Oxidative Stress in Vascular Pathophysiology)

Abstract

:
Elevated resistance of pulmonary circulation after chronic hypoxia exposure leads to pulmonary hypertension. Contributing to this pathological process is enhanced pulmonary vasoconstriction through both calcium-dependent and calcium sensitization mechanisms. Reactive oxygen species (ROS), as a result of increased enzymatic production and/or decreased scavenging, participate in augmentation of pulmonary arterial constriction by potentiating calcium influx as well as activation of myofilament sensitization, therefore mediating the development of pulmonary hypertension. Here, we review the effects of chronic hypoxia on sources of ROS within the pulmonary vasculature including NADPH oxidases, mitochondria, uncoupled endothelial nitric oxide synthase, xanthine oxidase, monoamine oxidases and dysfunctional superoxide dismutases. We also summarize the ROS-induced functional alterations of various Ca2+ and K+ channels involved in regulating Ca2+ influx, and of Rho kinase that is responsible for myofilament Ca2+ sensitivity. A variety of antioxidants have been shown to have beneficial therapeutic effects in animal models of pulmonary hypertension, supporting the role of ROS in the development of pulmonary hypertension. A better understanding of the mechanisms by which ROS enhance vasoconstriction will be useful in evaluating the efficacy of antioxidants for the treatment of pulmonary hypertension.

1. Introduction

The pulmonary circulation is normally a low resistance and low-pressure system with a mean pulmonary arterial pressure (mPAP) less than 20 mmHg [1]. Pulmonary hypertension (PH) is diagnosed by resting mPAP greater than 20 mmHg accompanied by pulmonary vascular resistance ≥ 3 Wood Units in pre-capillary PH, and classified as either idiopathic (pulmonary arterial hypertension, PAH, WHO Group 1) or secondary to left heart diseases (WHO Group 2), chronic hypoxic lung diseases (WHO Group 3), thrombosis (WHO Group 4), or of unclear reasons (WHO Group 5) [1].
PH is typically not diagnosed at early stages until the appearance of heart failure symptoms [2] such as dyspnea, palpitation and lower-extremity edema, which can ultimately lead to morbidity and mortality. This review will address many forms of PH with a focus on chronic hypoxia (CH)-induced PH, which occurs in patients with chronic obstructive pulmonary diseases (COPD), restrictive lung diseases, sleep apnea and in residents at high altitude.
Narrowing of pulmonary arteries (PAs) as a result of both structural (pulmonary arterial remodeling) and functional changes (vasoconstriction) contributes to increased vascular resistance that is pivotal to the pathogenesis of PH. Although pulmonary arterial wall thickening is observed in CH-induced PH [3,4,5], augmented vasoconstriction, manifested as both resting pulmonary arterial tone and reactivity to endogenous vasoconstrictors, plays an indispensable role in this disease [6,7,8,9,10]. Enhanced pulmonary arterial vasoconstriction results from pulmonary arterial smooth muscle cell (PASMC) hyperreactivity mediated by cytosolic Ca2+-dependent and Ca2+-sensitization mechanisms [11,12,13,14,15,16,17,18,19,20,21,22] as well as pulmonary arterial endothelial cell (PAEC) dysfunction via unbalanced production of vasoconstrictors over vasodilators [23]. Increased reactive oxygen species (ROS) have been widely reported to mediate augmented pulmonary arterial constriction [3,12,18,19,20,22,24,25,26,27,28] (Figure 1), and antioxidation strategies provide therapeutic efficacy in animal models of PH [3,29,30,31,32,33,34,35,36,37].

2. ROS in the Pathogenesis of PH

ROS are a group of oxygen-derived molecules with one or more unpaired electrons in their outer orbit. Superoxide anions (O2.−) are formed when molecular oxygen (O2) receives an electron, and is derived from various sources including NADPH oxidases (NOXs), mitochondria, endothelial nitric oxide synthase (eNOS), xanthine oxidase (XO) and monoamine oxidases (MAOs). The other two forms of ROS, peroxynitrite ion (ONOO) and hydrogen peroxide (H2O2), are derivatives of O2.−. Specifically, the combination of O2.− and nitric oxide (NO) produces ONOO and partial reduction of O2.− by superoxide dismutase (SOD) generates H2O2. There are three known SOD isoforms found in mammals including SOD1 (Cu-Zn SOD), located in the cytoplasm and intermembrane space of mitochondria, SOD2 (Mn SOD), located in the mitochondrial matrix, and SOD3 (Cu-Zn SOD), located extracellularly [38]. H2O2 is fully reduced to water (H2O) by catalase or glutathione peroxidase (Figure 2).
Under normal conditions, ROS are essential signaling molecules that are tightly regulated to maintain physiological homeostasis, regulate cellular proliferation, and host defense. Within the vasculature, ROS contribute to basal endothelial cell proliferation/migration [39,40], as well as smooth muscle cell differentiation [41]. ROS can also participate in vasomotor responses such as autoregulation [42], endothelium-dependent vasodilation [43,44], flow-mediated vasodilation [45], hypoxic pulmonary vasoconstriction (HPV) [46,47] and hyperoxia-induced vasoconstriction [48,49,50]. At physiological concentrations, H2O2 elicits vasodilation in the pulmonary circulation [51,52,53,54] and diminishes HPV in CH animals [55]. In addition, oxidation of protein kinase G Iα by oxidants (H2O2, glutathione disulfide, and protein-bound persulfides) following CH counteracts enhanced PA constriction and is protective during PH development [56], suggesting that under certain conditions ROS can play a protective role. However, H2O2 has also been reported to be detrimental in PH [29,57,58]. This contradictory effect of H2O2 in the pulmonary circulation may be due to a variety of conditions, including the concentration of H2O2, experimental setup, cell type affected and intracellular signaling mechanism involved. Other forms of ROS involved in the pathogenesis of PH include O2.− [3,12,26,29,59,60] and ONOO [61,62,63]. Although physiological levels of ROS are indispensable in maintaining vascular homeostasis, excess production leads to disease development [64,65,66,67] as detailed in the following sections.
O2.− is considered to be highly reactive and can mediate cell signaling either directly or through its derivatives, H2O2 and ONOO. These ROS participate in signaling transduction by making posttranslational modifications to ion channels, protein kinases and other signaling molecules [68,69,70]. Such protein modifications include oxidation of tyrosine, tryptophan, histidine, lysine, methionine and cysteine residues [70]. In this review, we discuss the ROS-induced functional alterations to relevant Ca2+ channels, K+ channels and other proteins that contribute to enhanced PA constriction in PH. Despite the observed effects of ROS, current knowledge about the exact chemical reactions, amino acid residues affected, and resulting protein structural changes is still limited. The most widely documented modifications in this setting are cysteine oxidative modifications (Figure 3), including S-glutathionylation (e.g., L-type voltage-gated Ca2+ channels [71,72] and STIM1 [73]), disulfide formation (e.g., ASIC1 [74], voltage-gated K+ channels [75,76] and RhoA [77]), and sulfenic acid formation (e.g., voltage-gated K+ channel [78]).

2.1. Increased ROS Production in PH

2.1.1. NADPH Oxidase Family

The NOX family consists of a group of enzymes that transfer an electron from NADPH to O2, therefore generating O2.−. NOX-derived ROS were first identified as effectors from phagocytes responsible for host defense [79] and later as mediators in various cellular processes. All enzymes within this family contain one of the seven ROS-generating catalytic homologs including NOX1, NOX2, NOX3, NOX4, NOX5, DUOX1 and DUOX2 [79]. Some of them are reported to be expressed within the pulmonary vasculature, as summarized in Table 1. NOX enzymes are comprised of several subunits, both catalytic and regulatory, that are located both intracellularly and extracellularly. This fact makes it possible for enzyme function to be regulated by its associated regulatory subunits as well as various intra- and extracellular signals [79,80].
NOX1 expression is greater in PAs from PAH patients compared to vessels from control patients [83] and contributes to the proliferation of both PAEC [83] and PASMC [86]. In monocrotaline-induced PAH model, NOX1 expression is increased in PASMCs [87]. Moreover, N-acetylcysteine, which suppresses NOX1 expression, is protective against monocrotaline-induced PAH [86]. In addition to PAH, NOX1 has also been shown to participate in PH elicited by CH as evidence by effects of genetic global deletion of NOX1 to abolish the CH-induced elevation in right ventricular systolic pressure (RVSP), right ventricle (RV) hypertrophy and PA remodeling in mice [98].
NOX2 expression is upregulated in response to prolonged [29] in vitro hypoxia. Using PAs isolated from wild type and gp91phox deficient mice, Liu et al. [24] discovered that NOX2-derived O2.− production in PAs is higher after CH exposure. Additionally, augmented PA contraction to ET-1 following CH is NOX2 dependent [24]. Evidence from our group also supports that NOX2-derived ROS contribute to enhanced PA constriction following CH [18,22].
NOX4 is unique among the NOX family since it is intrinsically active once expressed [99,100,101]. Although biochemical evidence suggests H2O2 is the major product of NOX4 [99,100,101], NOX4-dependent generation of both O2.− [82,97] and H2O2 [82,95,96] has been observed in the pulmonary vasculature. NOX4 expression in PAs from COPD patients [102] and CH mice [90] is higher than those from controls. NOX4 promotes proliferation of human PASMCs [90] and correlates with the severity of PA remodeling in COPD patients [102], suggesting a pathological role for NOX4 in CH-induced PH. In comparison to wild-type (WT) mice, Hood et al. [98] demonstrated that RVSP is diminished in NOX4 knockout (KO) mice following CH (10% O2 for 15 days). However, these NOX4 KO mice develop a similar degree of RV hypertrophy and PA remodeling as WT mice. In contrast, Veith et al. [103] reported that both global and inducible NOX4 KO mice exhibit similar elevations in RVSP after CH exposure (10% O2 for 21 days) as WT mice. The reason for these discrepant results is not clear, but may be due to differences in the duration of CH exposure or animal sex, as Hood et al. [98] studied only female mice. Since NOX4 does not account for monocrotaline-induced PAH [87], it appears that involvement of NOX4 in PH may differ depending on the model employed. In addition to evidence against a detrimental role of NOX4 in PH, a recent study demonstrates that increased disulfide protein kinase G Iα during CH, likely caused by NOX4-derived H2O2, opposes the pathogenesis of PH [56]. This possible beneficial effect of NOX4 in CH-induced PH is consistent with previous investigations showing the protective role of NOX4 in cardiovascular diseases. For example, endothelial NOX4 alleviates both angiotensin II-induced hypertension [104] and hemodynamic overload-induced cardiac remodeling [105].

2.1.2. Mitochondria

Mitochondria are double-membrane organelles responsible for efficient energy generation from the electron transport chain (ETC), which is located in the inner membrane. Electrons from NADH and FADH2, extracted by complex I (NADH dehydrogenase) and complex II (succinate dehydrogenase) respectively, flow through the ETC and are received by O2 at complex IV to generate H2O. A small amount of ROS are inevitably produced during this process because of “electron leak” [106]. A mitochondrial antioxidation system scavenges mitochondria-derived ROS (mitoROS) so that normal cell function can be maintained. This antioxidant system is comprised of SOD2 in the matrix and SOD1 in the intermembrane space, both of which convert O2.− into H2O2 [107,108]. H2O2 in the mitochondrial matrix is further detoxified by glutathione peroxidase 1(GPX1) [109,110] or catalase [29,110,111]. It is mainly reported that mitoROS are generated from complex I and III with O2.− as the primary product [112]. The role of mitoROS in CH-induced PH has been studied by several groups. Human PAECs exposed to prolonged hypoxia (72 h) have greater mitoROS levels versus normoxic controls [29]. MitoROS within PAEC participate in Ca2+ homeostasis as supported by data that higher intracellular Ca2+ in PAECs from SU5416/hypoxia-induced PAH rats versus those from normoxic animals is acutely diminished by the mitochondria-targeted antioxidant MitoQ [113]. A recent study from our laboratory similarly employed the mitochondrial antioxidants MitoQ and MitoTEMPO to demonstrate that mitoROS production is greater in PASMCs from CH neonatal rats compared to normotensive animals and contributes to enhanced PA vasoconstriction following CH [114]. Since MitoQ suppresses mitochondrial O2.− generation, it suggests the involvement of mitochondria-derived O2.− in the pathology of PH. Using genetically modified mice, other groups reported that mitochondrial H2O2 is also pathogenic in the development of PH elicited by CH. Evidence from Adesina et al. [29] show that indices of CH-induced PH, including RVSP, RV hypertrophy and PA remodeling, are attenuated by mitochondrial catalase overexpression, which breaks down mitochondrial H2O2, but are exacerbated by SOD2 overexpression, which increases mitochondrial H2O2.

2.1.3. Endothelial Nitric Oxide Synthase

NO, an important vasodilator in the pulmonary vasculature [115], is produced by eNOS using L-arginine as substrate, which requires the cofactor tetrahydrobiopterin (BH4) [116,117]. eNOS is constitutively expressed in endothelial cells and has 3 domains, a reductase domain in the C terminus, an oxygenase domain in the N terminus and a linking domain [118,119]. Binding of Ca2+/camodulin to the linking domain activates the enzyme, allowing NADPH oxidation to occur at the reductase domain. Flavin mononucleotide (FMN) and flavin adenine dinucleotide (FAD) of the reductase domain of one monomer pass electrons from NADPH to the heme-containing oxygenase domain of a second monomer via BH4 [117,119], which couples NADPH oxidation to L-arginine oxidation. NO is synthesized from O2 and L-arginine at the oxygenase domain through a multi-step chemical reaction including: (1) the combination of O2 with the heme group of the oxygenase domain to form a ferrous–dioxygen complex; (2) reduction of O2 within the ferrous-dioxygen complex to form H2O; and (3) oxidation of L-arginine to produce NO and L- citrulline [116,117,118,120]. Results from Vásquez-Vivar et al. [116] demonstrate that BH4 stabilizes the ferrous-dioxygen complex to prevent O2.− generation from eNOS. Thus, without adequate availability of BH4, eNOS produces O2.− [121]. O2.−, in turn, can oxidize BH4. BH4 oxidation is detrimental because it further reduces BH4 bioavailability and produces dihydrobiopterin (BH2) [122], a process that additionally favors O2.− generation from eNOS [123]. Interestingly, it is also reported that increases in BH2 alone, without a change of BH4 levels, are sufficient to induce O2.− generation [121], suggesting the BH4/BH2 ratio is key to regulation of eNOS function. Therefore, a deleterious cycle is established in which uncoupled eNOS produces O2.−, and O2.− further uncouples eNOS.
BH4-eNOS coupling is important in maintaining physiologically low pulmonary arterial pressure as evident by the effect of BH4 deficient mice to increase endothelial O2.− levels as well as right ventricular systolic pressure compared to wild type mice under normoxia, regardless of changes in eNOS expression [124]. In addition, pathological hypoxic exposure triggers more severe PH in BH4-deficient mice versus WT mice, which is attenuated by genetic BH4 restoration [124]. The importance of BH4-eNOS coupling in CH-induce PH is also supported by research findings from Dikalova et al. [31], in which oral BH4 administration attenuates the development of CH-induced PH in piglets. The therapeutic potential of BH4 is demonstrated by effects of BH4 treatment to reverse established CH-induced PH in rats [34]. Consistent with evidence that supplementation of BH4 [116,123,124,125,126] and an increased BH4/oxidized BH4 ratio [31,121,123] enhance eNOS activity, oral BH4 administration promotes eNOS activity, lowers lung O2.− levels, and reverses established CH-induced PH [34]. Additionally, BH4 is shown to be beneficial in the treatment of a rat PAH model [127].
In addition to BH4 oxidation, O2.− can mediate inhibition of eNOS by exogenous NO [63,128,129], a response that correlates with the clinical observation that sudden withdrawal of inhaled NO therapy worsens PH in children [130] and infants [131] with congenital heart diseases. In primary ovine PAEC cultures, Sheehy et al. [128] discovered that reduced eNOS activity by the NO donor sodium nitroprusside is partially restored by the O2.− scavenger Tiron. Since the reduction of eNOS activity by NO is not related to cell viability, eNOS expression, subcellular localization, or phosphorylation of eNOS [128], the mechanism by which ROS mediate NO-induced eNOS inhibition remains unclear in this cell model. More in-depth mechanisms are revealed by effects of NO inhalation to inhibit eNOS activity in a lamb model [63,129]. The involvement of ROS is implicated by the fact that O2.− and ONOO are increased in PAs following 24 h of NO inhalation [63]. The subsequent elevated nitration of eNOS by ONOO [63] is known to reduce enzyme activity [129]. Furthermore, lambs receiving polyethylene glycol-conjugated superoxide dismutase (PEG-SOD) at the same time of NO inhalation do not show rebound PH after acute NO withdrawal seen in those treated with vehicle [63]. Taken together, these data suggest exogenous NO leads to O2.− generation in the pulmonary vasculature, which reacts with NO to produce ONOO. The resultant nitration of eNOS suppresses its activity.

2.1.4. Xanthine Oxidase

The final two biochemical reactions of purine catabolism, namely conversions of hypoxanthine to xanthine to uric acid, are catalyzed by xanthine oxidoreductase (XOR). XOR consists of one molybdenum, two different iron-sulfur centers, and one FAD that all function as electron transporters [132,133,134]. An NAD+-dependent form of XOR, called xanthine dehydrogenase (XDH), is constitutively expressed, which fulfills purine degradation and generates NADH [135]. However, through oxidation of cysteine residues or proteolytic cleavage, XDH can be converted into xanthine oxidase (XO) [134,136]. Due to a decrease in NAD+ affinity and an increase in O2 affinity at the FAD site, XO exhibits high xanthine/O2 reductase activity instead of high xanthine/NAD+ reductase activity seen in XDH [132,133,134,135,136]. Therefore, purine catabolism catalyzed by XO produces ROS [132,133,134,135,136]. Experimental data from Kelley et al. [137] show that XO generates both H2O2 and O2.− with the former as the main product (>70% of ROS) under both normoxic (21% O2) and hypoxic conditions. Interestingly, when the O2 concertation is less than 10%, the proportion of H2O2 produced is inversely related to O2 concentration and can up to 90% in the presence of 1% O2 [137], indicating the involvement of H2O2 in XO-mediated diseases caused by hypoxia.
In the context of PH, enhanced XO activity upon hypoxic exposure has been confirmed in both in vitro [138] and in vivo [35,36] studies. Experimental inhibition of XO is protective against CH-induced increases in mPAP [35], RV hypertrophy [35,36] and pulmonary arterial wall thickening [35,36] in animal models. Consistently, compared to placebo treatment, XO inhibitor treatment (allopurinol) alleviates RV hypertrophy in COPD-associated PH patients with severe airflow limitation in a double-blinded randomized controlled clinical trial [139].

2.1.5. Monoamine Oxidases

Monoamine oxidases (MAOs) catalyze the oxidative deamination of bioactive amines and are found in brain as well as various human tissues [140]. MAO type A (MAO-A) and MAO type B (MAO-B) are two identified MAO isoforms characterized by different substrate preferences [141]. In particular, MAO-A oxidizes dopamine, norepinephrine and serotonin (5-HT), while MAO-B reacts with dopamine, phenylethylamine, benzylamine and tryptamine [142,143,144,145,146]. Substrate selectivity of MAOs is determined by phenylalanine residue 208 of MAO-A and isoleucine residue 199 of MAO-B [146]. During the oxidative deamination process, FAD, a cofactor of MAOs, delivers electrons from amines to molecular O2 to generate ROS [147], including O2.− [148,149] and H2O2 [150,151,152,153]. Therefore, MAOs can act as sources of ROS. One of the downstream targets of MAOs is mitochondria, which is consistent with evidence that MAOs are tethered to the outer membrane of mitochondria [154] via a C terminal transmembrane helix [155,156]. Specifically, MAOs mediate mitochondrial dysfunction [157,158,159,160] and promote mitoROS production [158,159]. Within the pulmonary circulation, MAO-A is expressed in PAs and contributes to O2.− generation triggered by 5-HT [148]. Preliminary observations from Sun and colleagues demonstrate that expression of MAO-A is upregulated in PAH patients [161] and that the MAO-A inhibitor clorgyline partially reverses indices of PH in a PAH rat model (SU5416/hypoxia), including RVSP, RV hypertrophy and PA remodeling [161,162]. However, the role of MAOs-derived ROS in vasoconstrictor responses of PAs and their contribution to CH-induced PH remains unclear.

2.2. Decreased Antioxidant Capacity in PH

Augmented ROS signaling can also result from decreased antioxidant capacity. Impaired SOD activity has been reported in a variety of PH models of animals and patients [33,163,164,165,166,167,168]. Aiming at rescuing the dysfunctional SOD system, SOD mimetics have been demonstrated to alleviate indices of PH following CH [3,32,37]. As mentioned before, three SOD isoforms are found in mammals [38] and all of them are reported to be important in the pathogenesis of PH.

2.2.1. SOD1

Expression of the predominant cytosolic SOD isoform, SOD1, is lower in PAs from CH piglets [165] and CH adult rats [28] in comparison to normoxic controls, a response associated with increased O2.− and decreased H2O2 levels. However, the role of SOD1 in CH-induced PH is not clear. Interestingly, compared to WT mice, SOD1 KO mice display elevated O2.− levels in PAs, exhibit enhanced vasoreactivity to ET-1, as well as greater RV hypertrophy, PA remodeling and greater RVSP under normoxia [166]. Collectively, these results indicate that loss of SOD1 in response to CH may contribute to the pathogenesis of PH.

2.2.2. SOD2

SOD2 is localized to the mitochondrial matrix [38]. SOD2 expression is reported to be downregulated in PH including in a CH mouse model [169], persistent PH lamb model [170], PAH patients [168] and a fawn-hooded rat model of PAH [164]. Loss of SOD2 in PASMCs during CH exposure may promote PA remodeling since SOD2 suppresses proliferation and promotes apoptosis of hypoxic cultures of human PASMCs [171]. It is also been shown that loss of SOD2 in PAECs is involved in elevated PA constriction in PH. In a PH neonatal lamb model established by ligation of the fetal patent ductus arteriosus during late gestation, SOD2 restoration in PAECs by adenovirus vectors reduces mitochondrial O2.− levels and restores eNOS expression [170], suggesting an improvement of PA dilation. Moreover, SOD2 transduction in PA rings from PH animals ameliorates their relaxation in response to the NO-dependent vasodilator, ATP, compared to control transduction [170]. Since the greater H2O2 production following restoration of SOD2 expression is thought to be responsible for the observed upregulation of eNOS [170], these findings suggest that mitochondria-derived H2O2 is protective. However, Adesina et al. [29] found that CH-induced RV hypertrophy, PA muscularization and increases in RSVP are exacerbated in a transgenic mouse model overexpressing SOD2 in comparison to WT mice. In this study, increased mitochondrial H2O2 is shown to be detrimental rather than protective.

2.2.3. SOD3

SOD3 locates extracellularly by binding to extracellular matrix components such as heparan sulfate proteoglycan, collagen and fibulin-5 [38]. Since introduction of extracellular ROS by administration of XO [172] and knockdown of SOD3 by siRNA [173] in cultured human PASMCs triggers pro-proliferative and anti-apoptotic phenotypic changes, it suggests extracellular O2.−, as well as SOD3 are likely important in the pathogenesis of PH. Considering that expression [28] and activity [28,173] of SOD3 in PAs are reduced by CH exposure, genetically modified animals with SOD3 deletion have been used to study its role in CH-induced PH development. Compared to control animals, mice with smooth muscle-specific SOD3 KO [174] or global SOD3 KO [175] exhibit greater RVSP, RV hypertrophy and pulmonary arterial wall thickening following CH. In line with this, SOD3 overexpression protects against CH-induced PH [30]. In addition to favoring PASMC proliferation, extracellular ROS also participate in CH-induced extracellular matrix remodeling as SMC SOD3 deletion augments CH-induced collagen deposition in PAs [174]. While dysfunctional SOD3 has been reported to be associated with PA remodeling, the role of SOD3 in enhanced PA vasoconstriction following CH is undetermined except for evidence that SOD3 helps to maintain normal eNOS function. Nozik-Grayck and colleagues [174] found that eNOS activation and GTP cyclohydrolase-1 (GTPCH-1, a key enzyme for BH4 synthesis) levels are diminished in lungs from smooth muscle-specific SOD3 KO mice exposed to CH, while eNOS expression is unaltered. These results are consistent with the notion that the loss of SOD3 contributes to development of PH.

3. ROS Modulation of Augmented PA Constriction

The effect of CH to augment PA vasoconstrictor reactivity has been convincingly demonstrated [6,7,8,9,10]. Smooth muscle contraction is triggered by an increase in intracellular Ca2+ levels ([Ca2+]i) via either Ca2+ influx or Ca2+ release from the sarcoplasmic reticulum (SR). Ca2+ binds to calmodulin and actives myosin light chain kinase (MLCK). When the regulatory light chain of myosin is phosphorylated by MLCK, cross-bridge cycling occurs and results in smooth muscle contraction (Ca2+-dependent mechanism). Contraction ends when phosphorylated myosin light chain is dephosphorylated by myosin light chain phosphatase (MLCP). Therefore, factors that inhibit MLCP activity can maintain smooth muscle contraction and contribute to prolonged vasoconstriction independent of changes in [Ca2+]i (Ca2+ sensitization mechanism). Both increases in [Ca2+]i in PASMCs [11,17,176,177,178,179,180] and Ca2+ sensitization [12,18,20,21,22,180,181,182] are known to mediate enhanced PA vasoconstriction in response to CH.

3.1. ROS Modulation of Ca2+-Dependent Vasoconstriction

3.1.1. Ca2+ Influx

Ca2+ influx is thought to contribute to the increase in [Ca2+]i in PASMCs after CH exposure, which can involve either voltage-gated calcium channels (VGCC) or non-selective cation channels (i.e., conduct both Ca2+ and Na+) including receptor-operated channels (ROC), store-operated channels (SOC), and mechanosensitive channels (MSCs) (Figure 4).
VGCC are gated by plasma membrane potential. Based on their sensitivity to depolarization, they are classified as high voltage-activated channels (L-, P/Q-, R-, N-type) and low voltage-activated (T-type) channels [183]. L-type and T-type VGCC are found in the pulmonary circulation [184]. Since membrane potential is a product of uneven distribution of Na+, K+ and Cl across the plasma membrane, the opening of non-selective cation channels (increased Na+ influx) and closing of K+ channels (reduced K+ efflux) lead to membrane depolarization and VGCC activation [185]. Plasma membrane depolarization is observed in PASMCs from CH animals [186,187] and PAH patients [188].
ROCs are controlled by diacylglycerol (DAG) generated from Gq protein-coupled receptor pathway activation [189]. SOCs are activated when intracellular SR Ca2+ stores are depleted [190]. When Ca2+ depletion in the SR is sensed by stromal interaction molecule (STIM) [190], STIM moves towards the plasma membrane and activates store-operated Ca2+ entry (SOCE) via Orai [191], acid-sensing ion channels (ASICs) [17,192] and transient receptor potential (TRP) channels [193].
Ca2+ channel expression and/or activity have been shown to be increased by CH and coupled to enhanced vasoconstriction (Table 2). Ca2+ signaling contributes to PA constriction [189]. However, it is worthwhile to note that some research findings may be animal model-specific. Previous data from our laboratory demonstrate the existence of differences in Ca2+ handling after CH exposure in two commonly used strains of rats [194]. In particular, CH induces an elevation in resting smooth muscle [Ca2+]i in Wistar rats but not in Sprague-Dawley (SD) rats [194]. Additionally, SOCE is attenuated by CH in SD rat while augmented in Wistar rats [194].

Voltage-Gated Calcium Channels

As summarized in Table 2, both positive and negative findings are documented for the role of VGCC in the pathogenesis of CH-induced PH. This discrepancy may be due to differences in animal species/strains and hypoxic protocols employed. Even though the role of VGCC in CH-induced PH is controversial, redox regulation of VGCC is possible. L-type VGCC can be inhibited by NO but stimulated by ONOO [210], a product from NO and O2.−. S-nitrosothiols are NO-donors that can cause either L-type VGCC inhibition [211] or activation [210]. Application of H2O2 [212,213,214] and oxidized glutathione [71,72] leads to Ca2+ influx through L-type VGCC [212,213,214]. Further study showed that of the Cav1.2 subunit of L-type VGCC is glutathionylated by H2O2 and oxidized glutathione (GSSG), which increases channel open probability and inward Ca2+ currents [71,72] (Figure 5). The well-known vasoconstrictor ET-1 can also stimulate L-type VGCC-mediated Ca2+ increases in PASMCs from CH Wistar rats [180,215]. Interestingly, this response is plasma membrane depolarization independent [180] but PKC and Rho kinase-dependent [215]. Redox modulation in this process is possible as both PKC [216] and Rho kinase [25] can be activated by oxidation. This possibility is further supported by the fact that ET-1 increases ROS production in PASMCs [22,217,218]. Even though this hypothesis is not tested in pulmonary circulation, stimulation of L-type VGCC by ET-1 in isolated cardiac myocytes is demonstrated to be O2.−-mediated [219].

Transient Receptor Potential Canonical 1 and 6 (TRPC 1 and 6) Channels

TRPC1 and TRPC6 contribute to increased [Ca2+]i in PASMCs as well as vasoconstriction following CH via involvement of SOCE and ROCE as indicated in Table 2. Their activities can be modulated by ROS. TRPC1 is mainly involved in SOCE in PASMC [178,200]. Administration of H2O2 increases STIM1/TRPC1 interactions and SOCE in cultured rat PASMCs [200]. TRPC6 is gated by diacylglycerol (DAG) [220,221,222,223]. The production of DAG can be facilitated by ROS. It is documented that NOX2-derived O2.− during ischemia-reperfusion and exogenous H2O2 phosphorylate and activate phospholipase C (PLC) [222] that generates DAG from cleavage of membrane PIP2. More direct evidence for ROS modulation of TRPC6 is that H2O2 triggers TRPC6-dependent Ca2+ influx in aortic vascular smooth muscle cells, as well as contraction in endothelium-denuded aorta [224] (Figure 5).

Transient Receptor Potential Vanilloid 4 (TRPV4) Channels

TRPV4, a member of the TRP channel superfamily, is a Ca2+ permeable non-selective cation channel. TRPV4 is expressed in all three layers of PAs, namely intimal (PAECs) [225,226,227], medial (PASMCs) [204,205] and adventitial (fibroblasts) layers [228]. Whereas TRPV4 expression in adventitial fibroblasts is upregulated by CH and contributes to excessive adventitial remodeling during the pathogenesis of PH [228], PASMC TRPV4 channels contribute to enhanced pulmonary vasoconstrictor reactivity following CH [204,205,206] (Table 2 for details).
In contrast, Ca2+ influx conducted by endothelial TRPV4 is coupled to vasodilation. For example, Ca2+ sparklets via endothelial TRPV4 activate eNOS to cause PA vasodilation [225,226,227]. Moreover, the PA vasodilator effect of a TRPV4 agonist is absent when NOS is inhibited [226,227,229], supporting the possibility that eNOS is downstream of TRPV4 in PAECs. Another possible mechanism underlying TRPV4-mediated PA dilation is through small/intermediate conductance Ca2+-activated K+ channels (SKCa/IKCa)-dependent endothelium-derived hyperpolarizing factor (EDHF) responses [226]. Intravenous injection of the TRPV4 agonist GSK101790A lowers pulmonary arterial pressure in normal rats [229], although its therapeutic potential in PH has not been documented.
Interestingly, TRPV4 activity in PAECs is enhanced by ROS. Extracellular H2O2 increases Ca2+ influx via TRPV4 in PAECs from mice and humans [230]. Mechanistically, this response requires TRPV4 phosphorylation by Fyn of the Src family kinases [230]. This process is facilitated by the scaffolding molecule CD36 that brings Fyn and TRPV4 together for efficient phosphorylation [231]. The possible phosphorylation site of TRPV4 that mediates its activation by H2O2 is serine 824 residue as demonstrated in human coronary artery endothelial cells channels [232]. Moreover, increased basal Ca2+ levels in PAECs from PAH rats are normalized by the SOD memetic TEMPOL, by mitochondria-targeted antioxidant MitoQ and by TRPV4 inhibitors [113], suggesting TRPV4 opening is maintained by endogenous ROS from mitochondria. This enhanced TRPV4-mediated Ca2+ entry in PAH animals contributes to proliferation and migration of PAECs [113]. Additionally, MitoQ attenuates TRPV4 agonist (GSK1016790A)-triggered Ca2+ influx observed in PAECs isolated from PAH rats [233] (Figure 5). Whether a similar ROS-induced TRPV4 activation mechanism in PASMCs contributes to elevated [Ca2+]i and vasoconstriction in PH [204,205,206], however, remains to be determined.

Acid-Sensing Ion Channel 1 (ASIC1)

ASICs are members of degenerin/epithelial sodium channels that are activated by extracellular protons. There are at least six known different ASIC subunits (ASIC1a, ASIC1b, ASIC2a, ASIC2b, ASIC3, and ASIC4) that exist in mammals and are encoded by 4 genes (ASIC1-4). Some ASICs, such as ASIC1a homomeric channels and ASIC1a/2b heteromeric channels, also have the ability to conduct Ca2+, therefore directly participating in intracellular Ca2+ homeostasis regulation [234,235]. ASIC subunits are cysteine-rich and modified by the cellular redox status. Reducing agents potentiate ASIC1 activity while oxidizing agents decrease ASIC1 current [74,236,237,238,239]. In addition, oxidizing agents like H2O2 introduce intersubunit disulfide bonds, thereby decreasing the amount of ASIC1a present on the cell surface and reduce acid-evoked currents [74]. Consistent with these studies, we found that H2O2 inhibited, and PEG-catalase augmented ASIC1-dependent Ca2+ influx in PASMCs [28]. Using a Wistar rat model of hypobaric hypoxia-induced PH, we found that PASMC O2.− levels are increased and H2O2 levels are decreased as a result of decreased SOD1 expression and activity [28]. This loss of endogenous H2O2 following CH contributes to the augmented ASIC1-dependent Ca2+ influx (Figure 5). The contribution of ASIC1 channels to CH-induced PH is summarized in Table 2.

Orai/STIM

STIM locates on SR. Translocation of STIM to plasma membrane upon SR depletion triggers SOCE via Orai channels. Orai/STIM participate in increases in resting cytosolic Ca2+ as well as SOCE following CH in PASMCs (Table 2). The resulting increase in intracellular Ca2+ levels is believed to couple to vasoconstriction but direct evidence for the contribution of Orai/STIM to PA constriction regulation is still absent. Orai/STIM-dependent SOCE can be regulated by ROS since oxidative stress upregulates STIM1 and Orai1 [200], increases STIM1/Orai1 interactions [200], facilitates STIM1 translocation to plasma membrane [73,240] and causes S-glutathionylation of cysteine 56 in STIM1 to trigger sustained Ca2+ entry that is independent of SR depletion [73] (Figure 5).

Mechanosensitive Channels (Aka Stretch-Activated Channels)

MSCs activated by plasma membrane stretch are permeable to Ca2+, therefore increasing [Ca2+]i. Ducret et al. [176] reported that MSC activity in PASMC is increased by CH exposure and contributes to myogenic tone of pulmonary arteries from CH rats whereas the PAs from normoxic animals do not exhibit tone (Table 2). Both O2.− and ONOO are shown to facilitate stretch-included activation of MSCs [241] (Figure 5).

3.1.2. K+ Efflux

K+ channels selectively conduct outward K+ currents that are important in maintaining physiological membrane potential. K+ channels can be classified into different categories depending on the gating mechanisms. Within the pulmonary vasculature, four different types of K+ channels have been identified, including voltage-gated K+ channels (KV), Ca2+-activated K+ channels (KCa), inwardly rectifying ATP-sensitive K+ channels (KATP) and four transmembrane segments-2 pore K+ channels (K2P) [242,243,244]. Loss of outward K+ currents can lead to membrane depolarization. Since plasma membrane depolarization is observed in PASMCs after CH exposure [186,187], a role for suppression of K+ channels in this response has been investigated. Membrane depolarization is an important PA constriction stimulus because it activates Ca2+ channels such as VGCCs. Evidence for the involvement of different K+ channels will be discussed below.
Loss of KV channel function is coupled to enhanced L-type VGCC activity and PA constriction. This is evident by data indicating that general KV channel blocker, 4-aminopyridine (4-AP), leads to L-type VGCC dependent increases in cytosolic Ca2+ as well as dose-dependent increases in basal pulmonary arterial tone [245]. Similar findings have been reported using blockers specific for KV7 (linopirdine and XE991) [246]. Therefore, KV channel downregulation and reduced KV currents observed in pulmonary vasculature [247,248,249,250,251,252,253] following CH are thought to mediate augmentation of PA constriction. This alteration is pathologically significant as restoration strategies have been shown to be beneficial in limiting PH [251,252,253]. KV1.5 and KV2.1 are of most importance in the context of CH-induced PH [247,248,249,250,251,252,253]. Detailed information is summarized in Table 3.
KV channels are redox-sensitive primary via the modification of cysteine residues. NOX4 is reported to be colocalized with KV1.5 in PASMCs and oxidizes cysteine residues in KV1.5 [249]. NOX4 inhibition alleviates the CH-induced reduction in KV currents [249]. Studies from Svoboda et al. [78] showed that ROS target the thiol (-SH) group of a cysteine (C581) residue at the C terminus of KV1.5, creating a sulfenic acid modification to KV1.5. Such modification results in a reduction of KV1.5 function by facilitating KV1.5 sequestration [78]. However, ROS modulation of KV2.1 within pulmonary circulation is understudied. In the central nervous system, ROS inhibit KV2.1 function by increasing its oligomerization [76,254]. Studies focused on molecular mechanisms of ROS-induced KV2.1 oligomerization revealed that KV2.1 oligomers are stabilized by disulfide bridges formed by oxidized cysteine residues at position 73 [75,76] and 710 [76].
Table 3. Suppression of KV channels following CH exposure contributes to PH.
Table 3. Suppression of KV channels following CH exposure contributes to PH.
Experimental ModelCH ProtocolDown-RegulationFunctional OutcomesRef.
Primary PASMCs from normal rats, unspecified strainHypoxia (3% O2, 5% CO2, and 92% N2), Normoxia (5% CO2 in air), 48–72 hKV1.2, KV1.5N/A[247]
Male Wistar ratsHypoxia (10±0.5 % O2), Normoxia (room air), 21 daysKV1.1, KV1.5, KV1.6, KV2.1, and KV4.3N/A[248]
Primary PASMCs from male SD ratsHypoxia (1% O2, 5% CO2, balance N2), Normoxia (5% CO2 in air), 48–72 hKV1.5 and KV2.1Decreased KV currents[249]
Primary PASMCs from normal SD ratsHypoxia (3% O2, 5% CO2, and 92% N2), Normoxia (5% CO2 in air), 60–72 hKV1.1, KV1.5, KV2.1, KV4.3, KV9.3Loss of channels causes decreased KV currents, membrane depolarization, increase cytosolic Ca2+[250]
Male Wistar ratsHypoxia (10% O2, balance N2), Normoxia (air), 21 daysKV1.5 and KV2.1 in PAsExpression restoration prevents elevation in mPAP, RV hypertrophy[251]
Male SD ratsHypoxia (10 % O2), Normoxia (room air), 14–21 daysKV1.5 and KV2.1 in PAsExpression restoration reverses and prevents PH indices following CH, including mPAP, PVR, RV hypertrophy, PA remodeling[252]
Male SD ratsHypoxia (380 mmHg), Normoxia (718 mmHg), 28 daysKV1.5 and KV2.1 in PAsExpression restoration (1) rescues CH-induced suppression of KV currents in PASMCs and (2) prevents RVSP elevation, RV hypertrophy and PA remodeling following CH[253]
Female wild-type control mice (C57BL/6XCBA strain)Hypoxia (10% O2), Normoxia (air), 14 daysN/AKV7 activator dilates pre-constricted PAs and prevents PH indices following CH, including mRVP, RV hypertrophy and PA remodeling[255]
Other K+ channels in pulmonary circulation, including KATP, KCa and K2P channels, are less well-studied in PH. Generally, these K+ channels exert a vasodilator effect in PAs when involved. The relevant findings are summarized in the following Table 4. Redox regulation of these K+ channels is possible but remains unclear in the pulmonary vasculature. As shown in Table 5, both activation and inhibition by ROS were observed in studies from other vascular beds.

3.2 ROS Participate in Ca2+ Sensitization

3.2.1 Rho Kinase Mediates Enhanced Ca2+ Sensitization

Enhanced vasoconstrictor responses in small PAs following CH can also be mediated by a Ca2+-independent mechanism in which vasoconstriction is independent of changes in intracellular Ca2+ levels. The contractile state of PASMCs results from the balance of MLCK and MLCP activities. Ca2+-independent vasoconstriction happens when phosphorylation of myosin light chain is maintained due to loss of MLCP activity. Phosphorylation of MLCP inhibits its function, which can be achieved by Rho kinase (ROK) either directly or indirectly via phosphorylated myosin light chain phosphatase inhibitor protein CPI-17 [281,282,283] (Figure 6). ROK is activated by GTP-bound RhoA [281,284]. Inhibition of ROK exerts a vasodilator effect on pulmonary vasculature [181,282,285] and therapeutic strategies targeting RhoA/ROK signaling is protective against CH-induced PH development in various animal models [282,286,287,288,289]. These observations suggest that RhoA/ROK represents a crucial mechanism underlying the pathogenesis of PH following CH. Pathophysiologically, ROK mediates the development of myogenic tone [19,21,181], along with enhanced vasoconstrictor reactivity to ET-1 [12,182] and membrane depolarizing stimuli following CH [18,19,20]. ROK can promote actin polymerization in PASMCs [19,288,290]. Our laboratory has demonstrated that such cytoskeletal remodeling actions of RhoA/ROK account for augmented PA constriction following CH [19]. ROK also facilities vasoconstriction by reducing eNOS expression as the inhibition of ROK increases eNOS expression in lungs of CH mice [282].

3.2.2 ROS Regulation of RhoA/ROK Signaling

Generally, the RhoA/ROK pathway can be activated by hypoxia [290], by plasma membrane depolarization [20,291,292,293,294] and by signals from G protein-coupled receptors, receptor tyrosine kinases, cytokine receptors and integrins [281,284]. Activated ROK phosphorylates the myosin phosphatase target subunit 1 (MYPT1) of MLCP at multiple threonine and serine residues [295], therefore inhibiting MLCP. Direct evidence for redox regulation of ROK activity is that the O2.− donor, LY83583, increases ROK-dependent MYPT1 phosphorylation in PAs [27]. Additionally, ROS can mediate ROK activation in PAs in response to stimuli such as U46619 [25] and CH [12,20]. ROS are important in linking various pathogenic stimuli, such as receptor activation and membrane depolarization, to ROK-dependent Ca2+ sensitization in PASMCs in the setting of CH, therefore contributing to enhanced vasoconstriction [9].
The molecular mechanism by which ROS modulate RhoA/ROK signaling is not fully understood. Upregulation of RhoA under hypoxia appears to be downstream of ROS production [296]. RhoA is a member of Rho GTPase family whose function is regulated by guanine nucleotide-binding state [297]. RhoA is activated when binds to GTP with facilitation from guanine nucleotide exchange factors (GEFs) and deactivated by hydrolysis of GTP [297]. ROS have been shown to activate RhoA by targeting its redox-sensitive GXXXXGK(S/T)C motif, which determines guanine nucleotide dissociation [77,298]. Within this motif, cysteine residues 16 and 20 are critical for ROS modulation of RhoA activity [77,299].
Previous work from our laboratory has focused on delineating the contribution of ROS-dependent myofilament Ca2+ sensitization to vasoreactivity following CH [12,18,19,20,21,22]. PAs from CH rats have greater tone compared to vessels from control animals without a difference in [Ca2+]i in PASMCs [19,21], suggesting the importance of the Ca2+ sensitization mechanism in maintaining elevated basal contractile state of PAs following CH. CH exposure also augments vasoconstriction to agonists independent of changes in Ca2+ because ET-1 [12] and membrane depolarization (KCl) [20] trigger greater constriction in Ca2+-permeabilized PAs from CH rats versus normoxic controls. Such differences are abolished by ROS scavengers alone [12,20], ROK inhibition alone [12,20] or combination of ROS scavenger and ROK [20]. Considering RhoA activation upon ET-1 [12] and KCl [20] stimulation in PAs requires ROS, it suggests ROS signal through ROK to facilitate CH-induced augmentation of myofilament Ca2+ sensitization. Furthermore, elevated KCl-induced vasoconstriction in Ca2+-clamped PAs from CH rats is normalized by NOX2 inhibition [18], indicating NOX2 as the enzymatic source of ROS involved in Ca2+ sensitization regulation. Upstream of NOX2 is epidermal growth factor receptor (EGFR) activation as EGFR is activated by KCl and contributes to KCl-induced ROS generation from NOX2 [18]. In summary, an EGFR-NOX2-ROK-mediated Ca2+ sensitization mechanism mediates CH-induced augmentation of PA vasoconstriction (Figure 6). This signaling is pathologically important because it participates in the development of PH following CH [22].

4. Conclusion/Perspective

This review summarizes our understanding of ROS in enhanced PA constriction in the disease of PH, with an emphasis on CH-associated PH. Pathological ROS signaling following CH is the outcome of increased production from various enzymatic sources, as well as dysfunctional scavenging systems. ROS participate in PA vasomotor tone regulation following CH by modulating Ca2+ influx, K+ efflux and myofilament Ca2+ sensitization. Although ROS have convincingly reported to have an effect on these processes, little is known about the precise ROS-induced modifications to the relevant ion channels and signaling molecules. Limited studies indicate that ROS cause protein phosphorylation at serine residues and target cysteine residues to introduce modifications such as S-glutathionylation, disulfide formation and sulfenic acid formation. However, there are some potential pitfalls associated with such studies. First, the majority of current knowledge about ROS regulation of ion channels is gathered from experiments applying exogenous/extracellular oxidizing reagents, which fail to fully reflect the actions of intracellular ROS seen in PH. Second, there is a lack of pulmonary circulation-specific evidence regarding redox modifications. Third, it remains unclear how these identified ROS-induced modifications alter functions of the affected ion channels and signal transducers. Future studies are therefore needed to address these limitations. Furthermore, the upstream regulatory mechanisms that regulate enzymatic sources of ROS in PH are not fully understood. Such knowledge will be valuable in designing therapies specific to the disease while preserving the physiological functions of ROS.
Considering the importance of ROS in mediating increases in pulmonary vascular resistance (PVR), a number of groups have attempted to develop novel therapeutic strategies for PH/PAH by preventing ROS production or scavenging ROS. Pre-clinical studies in PH/PAH animal models have shown promising results [3,29,30,31,32,33,34,35,36,37,300,301]. It is likely that oxidative signaling is also involved in human PH/PAH as oxidative stress is increased in chronic high-altitude residents [302] and in PH patients [303]. Moreover, in PAH patients, oxidative stress markers correlate with an adverse prognosis [304], and reducing oxidative stress by epoprostenol [305] or by recombinant human angiotensin converting enzyme type 2 (rhACE2) [306] is associated with decreases in PVR in small scale clinical trials. Aside from PVR, preserving or improving right heart function is a crucial goal for disease management [307] because cor pulmonale occurring during PH is fatal. Unfortunately, antioxidant therapy has shown little success in cardioprotection to date. Oral administration of antioxidant coenzyme Q (CoQ) improves left and right heart functions in PAH patients evaluated by echocardiography [308]. Additionally, cardiac magnetic resonance imaging demonstrates that the XO inhibitor, allopurinol, alleviates right ventricular hypertrophy in COPD-associated PH patients with severe airflow limitation [139]. However, in both trials, standard clinical cardiac function biomarkers, such as 6-minite walk distance and brain natriuretic peptide (BNP) levels, are not affected [139,308].
To our knowledge, these are the only available clinical trials [139,305,306,308] aimed at addressing the therapeutic potential of antioxidation strategies in PH/PAH to now. It is important to note that epoprostenol [305] and rhACE2 [306] used in clinical trials do not act primarily as antioxidants, and other classical antioxidants including SOD memetic and SOD/catalase mimetic have not been studied yet. Also, these pilot clinical trials have relatively small cohorts and fail to provide clinical details about optimal dose, treatment protocol, side effects and population generalizability. Therefore, multicenter double-blinded randomized controlled clinical trials are needed before drawing a firm conclusion about the therapeutic value of antioxidants in pulmonary hypertension.

Author Contributions

Writing—original draft preparation, S.Y.; writing—review and editing, S.Y., T.C.R. and N.L.J.; visualization, S.Y.; supervision, T.C.R. and N.L.J.; funding acquisition, T.C.R. and N.L.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by R01 HL 132883 to T.C.R. and R01 HL 111084 to N.L.J.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

[Ca2+]iIntracellular Ca2+ level
5-HTSerotonin
ASICAcid-sensing ion channel
BH2Dihydrobiopterin
BH4Tetrahydrobiopterin
BKCaLarge conductance Ca2+-activated K+ channels
BNPBrain natriuretic peptide
Ca2+Calcium
CHChronic hypoxia
COPDChronic obstructive pulmonary disease
CoQCoenzyme Q
DAGDiacylglycerol
DTNB5,5′-Dithiobis(2-nitrobenzoic acid)
DTTDithiothreitol
ECEndothelial cell
EGFREpidermal growth factor receptor
eNOSEndothelial nitric oxide synthase
ET-1Endothelin-1
ETCElectron transport chain
FADFlavin adenine dinucleotide
GPCRG protein coupled receptor
GSHReduced glutathione
GSSGOxidized glutathione
H2O2Hydrogen peroxide
HPVHypoxic pulmonary vasoconstriction
IKCaIntermediate conductance Ca2+-activated K+ channels
IP3Inositol triphosphate
K+Potassium
K2PFour transmembrane segments-2 pores K+ channels
KATPATP-sensitive K+ channels
KCaCa2+-activated K+ channels
KOKnockout
KVVoltage-gated K+ channels
MAOMonoamine oxidase
MitoROSMitochondria-derived ROS
MLCKMyosin light chain kinase
MLCPMyosin light chain phosphatase
mPAPMean pulmonary arterial pressure
MSC Mechanosensitive channel
MYPT1Myosin phosphatase target subunit 1
Na+Sodium
NADPHReduced nicotinamide adenine dinucleotide phosphate
NONitric oxide
NOXNADPH oxidase
O2.−Superoxide anion
ONOOPeroxynitrite ion
PAPulmonary artery
PAECPulmonary arterial endothelial cell
PAHPulmonary arterial hypertension
PASMCPulmonary arterial smooth muscle cell
PEG-SODPolyethylene glycol-conjugated SOD
PHPulmonary hypertension
PIP2Phosphatidylinositol 4,5-bisphosphate
PLCPhospholipase C
PVRPulmonary vascular resistance
rhACE2Recombinant human angiotensin converting enzyme type 2
ROCReceptor-operated channel
ROCEReceptor-operated Ca2+ entry
ROKRho kinase
ROSReactive oxygen species
RV Right ventricular
RVSPRight ventricular systolic pressure
SKCaSmall conductance Ca2+-activated K+ channels
SMCSmooth muscle cell
SOCStore-operated channel
SOCEStore-operated Ca2+ entry
SODSuperoxide dismutase
SRSarcoplasmic reticulum
STIMStromal interaction molecule
TRPCTransient receptor potential canonical
TRPV4Transient receptor potential vanilloid 4
VGCCVoltage-gated calcium channel
WHOWorld health organization
WTWild-type
XDHXanthine dehydrogenase
XOXanthine oxidase
XORXanthine oxidoreductase

References

  1. Simonneau, G.; Montani, D.; Celermajer, D.S.; Denton, C.P.; Gatzoulis, M.A.; Krowka, M.; Williams, P.G.; Souza, R. Haemodynamic definitions and updated clinical classification of pulmonary hypertension. Eur. Respir. J. 2019, 53, 1801913. [Google Scholar] [CrossRef]
  2. Brown, L.M.; Chen, H.; Halpern, S.; Taichman, D.; McGoon, M.D.; Farber, H.W.; Frost, A.E.; Liou, T.G.; Turner, M.; Feldkircher, K.; et al. Delay in recognition of pulmonary arterial hypertension: Factors identified from the REVEAL Registry. Chest 2011, 140, 19–26. [Google Scholar] [CrossRef] [Green Version]
  3. Jernigan, N.L.; Naik, J.S.; Weise-Cross, L.; Detweiler, N.D.; Herbert, L.M.; Yellowhair, T.R.; Resta, T.C. Contribution of reactive oxygen species to the pathogenesis of pulmonary arterial hypertension. PLoS ONE 2017, 12, e0180455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Stenmark, K.R.; Fagan, K.A.; Frid, M.G. Hypoxia-Induced Pulmonary Vascular Remodeling. Circ. Res. 2006, 99, 675–691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Young, J.M.; Williams, D.R.; Thompson, A.A.R. Thin Air, Thick Vessels: Historical and Current Perspectives on Hypoxic Pulmonary Hypertension. Front. Med. (Lausanne) 2019, 6, 93. [Google Scholar] [CrossRef] [Green Version]
  6. Cahill, E.; Rowan, S.C.; Sands, M.; Banahan, M.; Ryan, D.; Howell, K.; McLoughlin, P. The pathophysiological basis of chronic hypoxic pulmonary hypertension in the mouse: Vasoconstrictor and structural mechanisms contribute equally. Exp. Physiol. 2012, 97, 796–806. [Google Scholar] [CrossRef] [PubMed]
  7. Shimoda, L.A.; Sham, J.S.; Sylvester, J.T. Altered pulmonary vasoreactivity in the chronically hypoxic lung. Physiol. Res. 2000, 49, 549–560. [Google Scholar] [PubMed]
  8. Jernigan, N.L.; Resta, T.C. Calcium Homeostasis and Sensitization in Pulmonary Arterial Smooth Muscle. Microcirculation 2014, 21, 259–271. [Google Scholar] [CrossRef] [PubMed]
  9. Resta, T.C.; Broughton, B.R.S.; Jernigan, N.L. Reactive oxygen species and RhoA signaling in vascular smooth muscle: Role in chronic hypoxia-induced pulmonary hypertension. Adv. Exp. Med. Biol. 2010, 661, 355–373. [Google Scholar] [CrossRef]
  10. Stenmark, K.R.; McMurtry, I.F. Vascular remodeling versus vasoconstriction in chronic hypoxic pulmonary hypertension: A time for reappraisal? Circ. Res. 2005, 97, 95–98. [Google Scholar] [CrossRef] [Green Version]
  11. Undem, C.; Luke, T.; Shimoda, L.A. Contribution of elevated intracellular calcium to pulmonary arterial myocyte alkalinization during chronic hypoxia. Pulm. Circ. 2016, 6, 93–102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Jernigan, N.L.; Walker, B.R.; Resta, T.C. Reactive oxygen species mediate RhoA/Rho kinase-induced Ca2+ sensitization in pulmonary vascular smooth muscle following chronic hypoxia. Am. J. Physiol. Lung Cell. Mol. Physiol. 2008, 295, L515–L529. [Google Scholar] [CrossRef] [PubMed]
  13. Gilbert, G.; Ducret, T.; Marthan, R.; Savineau, J.-P.; Quignard, J.-F. Stretch-induced Ca2+ signalling in vascular smooth muscle cells depends on Ca2+ store segregation. Cardiovasc. Res. 2014, 103, 313–323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Wang, J.; Xu, C.; Zheng, Q.; Yang, K.; Lai, N.; Wang, T.; Wang, J.; Lu, W. Orai1, 2, 3 and STIM1 promote store-operated calcium entry in pulmonary arterial smooth muscle cells. Cell Death Discov. 2017, 3, 17074. [Google Scholar] [CrossRef] [Green Version]
  15. Jernigan, N.L.; Paffett, M.L.; Walker, B.R.; Resta, T.C. ASIC1 contributes to pulmonary vascular smooth muscle store-operated Ca2+ entry. Am. J. Physiol. Lung Cell. Mol. Physiol. 2009, 297, L271–L285. [Google Scholar] [CrossRef] [Green Version]
  16. Herbert, L.M.; Resta, T.C.; Jernigan, N.L. RhoA increases ASIC1a plasma membrane localization and calcium influx in pulmonary arterial smooth muscle cells following chronic hypoxia. Am. J. Physiol. Cell. Physiol. 2018, 314, C166–C176. [Google Scholar] [CrossRef]
  17. Jernigan, N.L.; Herbert, L.M.; Walker, B.R.; Resta, T.C. Chronic hypoxia upregulates pulmonary arterial ASIC1: A novel mechanism of enhanced store-operated Ca2+ entry and receptor-dependent vasoconstriction. Am. J. Physiol. Cell. Physiol. 2012, 302, C931–C940. [Google Scholar] [CrossRef]
  18. Norton, C.E.; Broughton, B.R.; Jernigan, N.L.; Walker, B.R.; Resta, T.C. Enhanced depolarization-induced pulmonary vasoconstriction following chronic hypoxia requires EGFR-dependent activation of NAD(P)H oxidase 2. Antioxid. Redox Signal. 2013, 18, 1777–1788. [Google Scholar] [CrossRef] [Green Version]
  19. Weise-Cross, L.; Sands, M.A.; Sheak, J.R.; Broughton, B.R.S.; Snow, J.B.; Bosc, L.V.G.; Jernigan, N.L.; Walker, B.R.; Resta, T.C. Actin polymerization contributes to enhanced pulmonary vasoconstrictor reactivity after chronic hypoxia. Am. J. Physiol. Heart Circ. Physiol. 2018, 314, H1011–H1021. [Google Scholar] [CrossRef]
  20. Broughton, B.R.S.; Jernigan, N.L.; Norton, C.E.; Walker, B.R.; Resta, T.C. Chronic hypoxia augments depolarization-induced Ca2+ sensitization in pulmonary vascular smooth muscle through superoxide-dependent stimulation of RhoA. Am. J. Physiol. Lung Cell. Mol. Physiol. 2010, 298, L232–L242. [Google Scholar] [CrossRef] [Green Version]
  21. Broughton, B.R.S.; Walker, B.R.; Resta, T.C. Chronic hypoxia induces Rho kinase-dependent myogenic tone in small pulmonary arteries. Am. J. Physiol. Lung Cell. Mol. Physiol. 2008, 294, L797–L806. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Norton, C.E.; Sheak, J.R.; Yan, S.; Weise-Cross, L.; Jernigan, N.L.; Walker, B.R.; Resta, T.C. Augmented Pulmonary Vasoconstrictor Reactivity after Chronic Hypoxia Requires Src Kinase and Epidermal Growth Factor Receptor Signaling. Am. J. Respir. Cell Mol. Biol. 2020, 62, 61–73. [Google Scholar] [CrossRef]
  23. Budhiraja, R.; Tuder, R.M.; Hassoun, P.M. Endothelial Dysfunction in Pulmonary Hypertension. Circulation 2004, 109, 159–165. [Google Scholar] [CrossRef]
  24. Liu, J.Q.; Erbynn, E.M.; Folz, R.J. Chronic hypoxia-enhanced murine pulmonary vasoconstriction: Role of superoxide and gp91phox. Chest 2005, 128 (Suppl. S6), 594S–596S. [Google Scholar] [CrossRef]
  25. Mackay, C.E.; Shaifta, Y.; Snetkov, V.V.; Francois, A.A.; Ward, J.P.; A Knock, G. ROS-dependent activation of RhoA/Rho-kinase in pulmonary artery: Role of Src-family kinases and ARHGEF1. Free Radic. Biol. Med. 2017, 110, 316–331. [Google Scholar] [CrossRef] [Green Version]
  26. Jernigan, N.L.; Walker, B.R.; Resta, T.C. Endothelium-derived reactive oxygen species and endothelin-1 attenuate NO-dependent pulmonary vasodilation following chronic hypoxia. Am. J. Physiol. Lung Cell. Mol. Physiol. 2004, 287, L801–L808. [Google Scholar] [CrossRef] [Green Version]
  27. Knock, G.A.; Snetkov, V.A.; Shaifta, Y.; Connolly, M.; Drndarski, S.; Noah, A.; Pourmahram, G.E.; Becker, S.; Aaronson, P.I.; Ward, J.P. Superoxide constricts rat pulmonary arteries via Rho-kinase-mediated Ca(2+) sensitization. Free Radic. Biol. Med. 2009, 46, 633–642. [Google Scholar] [CrossRef] [PubMed]
  28. Plomaritas, D.R.; Herbert, L.M.; Yellowhair, T.R.; Resta, T.C.; Bosc, L.V.G.; Walker, B.R.; Jernigan, N.L. Chronic hypoxia limits H2O2-induced inhibition of ASIC1-dependent store-operated calcium entry in pulmonary arterial smooth muscle. Am. J. Physiol. Lung Cell. Mol. Physiol. 2014, 307, 419–430. [Google Scholar] [CrossRef] [Green Version]
  29. Adesina, S.E.; Kang, B.-Y.; Bijli, K.M.; Ma, J.; Cheng, J.; Murphy, T.C.; Hart, C.M.; Sutliff, R.L. Targeting mitochondrial reactive oxygen species to modulate hypoxia-induced pulmonary hypertension. Free Radic. Biol. Med. 2015, 87, 36–47. [Google Scholar] [CrossRef] [Green Version]
  30. Ahmed, M.N.; Zhang, Y.; Codipilly, C.; Zaghloul, N.; Patel, D.; Wolin, M.; Miller, E.J. Extracellular Superoxide Dismutase Overexpression Can Reverse the Course of Hypoxia-Induced Pulmonary Hypertension. Mol. Med. 2012, 18, 38–46. [Google Scholar] [CrossRef] [PubMed]
  31. Dikalova, A.; Aschner, J.L.; Kaplowitz, M.R.; Summar, M.; Fike, C.D. Tetrahydrobiopterin oral therapy recouples eNOS and ameliorates chronic hypoxia-induced pulmonary hypertension in newborn pigs. Am. J. Physiol. Lung Cell. Mol. Physiol. 2016, 311, L743–L753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Elmedal, B.; De Dam, M.Y.; Mulvany, M.J.; Simonsen, U. The superoxide dismutase mimetic, tempol, blunts right ventricular hypertrophy in chronic hypoxic rats. Br. J. Pharmacol. 2004, 141, 105–113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Fike, C.D.; Dikalova, A.; Slaughter, J.C.; Kaplowitz, M.; Zhang, Y.; Aschner, J.L. Reactive Oxygen Species-Reducing Strategies Improve Pulmonary Arterial Responses to Nitric Oxide in Piglets with Chronic Hypoxia-Induced Pulmonary Hypertension. Antioxid. Redox Signal. 2013, 18, 1727–1738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Francis, B.N.; Hale, A.; Channon, K.M.; Wilkins, M.R.; Zhao, L. Effects of tetrahydrobiopterin oral treatment in hypoxia-induced pulmonary hypertension in rat. Pulm. Circ. 2014, 4, 462–470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Hoshikawa, Y.; Ono, S.; Suzuki, S.; Tanita, T.; Chida, M.; Song, C.; Noda, M.; Tabata, T.; Voelkel, N.F.; Fujimura, S. Generation of oxidative stress contributes to the development of pulmonary hypertension induced by hypoxia. J. Appl. Physiol. 2001, 90, 1299–1306. [Google Scholar] [CrossRef] [Green Version]
  36. Jankov, R.P.; Kantores, C.; Pan, J.; Belik, J. Contribution of xanthine oxidase-derived superoxide to chronic hypoxic pulmonary hypertension in neonatal rats. Am. J. Physiol. Lung Cell. Mol. Physiol. 2008, 294, L233–L245. [Google Scholar] [CrossRef]
  37. Villegas, L.R.; Kluck, D.; Field, C.; Oberley-Deegan, R.E.; Woods, C.; Yeager, M.E.; El Kasmi, K.C.; Savani, R.C.; Bowler, R.P.; Nozik-Grayck, E. Superoxide Dismutase Mimetic, MnTE-2-PyP, Attenuates Chronic Hypoxia-Induced Pulmonary Hypertension, Pulmonary Vascular Remodeling, and Activation of the NALP3 Inflammasome. Antioxid. Redox Signal. 2013, 18, 1753–1764. [Google Scholar] [CrossRef] [Green Version]
  38. Fukai, T.; Ushio-Fukai, M. Superoxide Dismutases: Role in Redox Signaling, Vascular Function, and Diseases. Antioxid. Redox Signal. 2011, 15, 1583–1606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Petry, A.; Djordjevic, T.; Weitnauer, M.; Kietzmann, T.; Hess, J.; Görlach, A. NOX2 and NOX4 Mediate Proliferative Response in Endothelial Cells. Antioxid. Redox Signal. 2006, 8, 1473–1484. [Google Scholar] [CrossRef]
  40. Abid, M.R.; Kachra, Z.; Spokes, K.C.; Aird, W.C. NADPH oxidase activity is required for endothelial cell proliferation and migration. FEBS Lett. 2000, 486, 252–256. [Google Scholar] [CrossRef] [Green Version]
  41. Clempus, R.E.; Sorescu, D.; Dikalova, A.E.; Pounkova, L.; Jo, P.; Sorescu, G.P.; Schmidt, H.H.H.; Lassègue, B.; Griendling, K.K. Nox4 is required for maintenance of the differentiated vascular smooth muscle cell phenotype. Arter. Thromb. Vasc. Biol. 2007, 27, 42–48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Yada, T. Hydrogen Peroxide, an Endogenous Endothelium-Derived Hyperpolarizing Factor, Plays an Important Role in Coronary Autoregulation In Vivo. Circulation 2003, 107, 1040–1045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Sindler, A.L.; Reyes, R.; Chen, B.; Ghosh, P.; Gurovich, A.N.; Kang, L.S.; Cardounel, A.J.; Delp, M.D.; Muller-Delp, J. Age and exercise training alter signaling through reactive oxygen species in the endothelium of skeletal muscle arterioles. J. Appl. Physiol. 2013, 114, 681–693. [Google Scholar] [CrossRef]
  44. Muñoz, M.; Martínez, M.P.; López-Oliva, M.E.; Rodríguez, C.; Corbacho, C.; Carballido, J.; García-Sacristán, A.; Hernández, M.; Rivera, L.; Medina, J.S.; et al. Hydrogen peroxide derived from NADPH oxidase 4- and 2 contributes to the endothelium-dependent vasodilatation of intrarenal arteries. Redox Biol. 2018, 19, 92–104. [Google Scholar] [CrossRef] [PubMed]
  45. Drouin, A.; Thorin, E. Flow-induced dilation is mediated by Akt-dependent activation of endothelial nitric oxide synthase-derived hydrogen peroxide in mouse cerebral arteries. Stroke 2009, 40, 1827–1833. [Google Scholar] [CrossRef] [Green Version]
  46. Ahmad, M.; Kelly, M.R.; Zhao, X.; Kandhi, S.; Wolin, M.S. Roles for Nox4 in the contractile response of bovine pulmonary arteries to hypoxia. Am. J. Physiol. Heart Circ. Physiol. 2010, 298, H1879–H1888. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Murtaza, G.; Paddenberg, R.; Pfeil, U.; Goldenberg, A.; Mermer, P.; Kummer, W. Hypoxia-induced pulmonary vasoconstriction of intra-acinar arteries is impaired in NADPH oxidase 4 gene-deficient mice. Pulm. Circ. 2018, 8, 2045894018808240. [Google Scholar] [CrossRef] [Green Version]
  48. Mak, S.; Egri, Z.; Tanna, G.; Colman, R.; Newton, G.E. Vitamin C prevents hyperoxia-mediated vasoconstriction and impairment of endothelium-dependent vasodilation. Am. J. Physiol. Heart Circ. Physiol. 2002, 282, H2414–H2421. [Google Scholar] [CrossRef] [Green Version]
  49. Rousseau, A.; Tesselaar, E.; Henricson, J.; Sjöberg, F. Prostaglandins and Radical Oxygen Species Are Involved in Microvascular Effects of Hyperoxia. J. Vasc. Res. 2010, 47, 441–450. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Zhilyaev, S.Y.; Moskvin, A.N.; Platonova, T.F.; Gutsaeva, D.R.; Churilina, I.V.; Demchenko, I.T. Hyperoxic vasoconstriction in the brain is mediated by inactivation of nitric oxide by superoxide anions. Neurosci. Behav. Physiol. 2003, 33, 783–787. [Google Scholar] [CrossRef]
  51. Burke, T.M.; Wolin, M.S. Hydrogen peroxide elicits pulmonary arterial relaxation and guanylate cyclase activation. Am. J. Physiol. Heart Circ. Physiol. 1987, 252 Pt 2, H721–H732. [Google Scholar] [CrossRef]
  52. Michelakis, E.D.; Hampl, V.; Nsair, A.; Wu, X.; Harry, G.; Haromy, A.; Gurtu, R.; Archer, S.L. Diversity in Mitochondrial Function Explains Differences in Vascular Oxygen Sensing. Circ. Res. 2002, 90, 1307–1315. [Google Scholar] [CrossRef]
  53. Neo, B.H.; Patel, D.; Kandhi, S.; Wolin, M.S. Roles for cytosolic NADPH redox in regulating pulmonary artery relaxation by thiol oxidation-elicited subunit dimerization of protein kinase G1α. Am. J. Physiol. Heart Circ. Physiol. 2013, 305, H330–H343. [Google Scholar] [CrossRef] [Green Version]
  54. Neo, B.H.; Kandhi, S.; Wolin, M.S. Roles for soluble guanylate cyclase and a thiol oxidation-elicited subunit dimerization of protein kinase G in pulmonary artery relaxation to hydrogen peroxide. Am. J. Physiol. Heart Circ. Physiol. 2010, 299, H1235–H1241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Patel, D.; Alhawaj, R.; Wolin, M.S. Exposure of mice to chronic hypoxia attenuates pulmonary arterial contractile responses to acute hypoxia by increases in extracellular hydrogen peroxide. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2014, 307, R426–R433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Rudyk, O.; Rowan, A.; Prysyazhna, O.; Krasemann, S.; Hartmann, K.; Zhang, M.; Shah, A.M.; Ruppert, C.; Weiss, A.; Schermuly, R.T.; et al. Oxidation of PKGIα mediates an endogenous adaptation to pulmonary hypertension. Proc. Natl. Acad. Sci. USA 2019, 116, 13016–13025. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Wedgwood, S.; Lakshminrusimha, S.; Fukai, T.; Russell, J.A.; Schumacker, P.T.; Steinhorn, R. Hydrogen Peroxide Regulates Extracellular Superoxide Dismutase Activity and Expression in Neonatal Pulmonary Hypertension. Antioxid. Redox Signal. 2011, 15, 1497–1506. [Google Scholar] [CrossRef] [PubMed]
  58. Wedgwood, S.; Steinhorn, R.; Bunderson, M.; Wilham, J.; Lakshminrusimha, S.; Brennan, L.A.; Black, S.M. Increased hydrogen peroxide downregulates soluble guanylate cyclase in the lungs of lambs with persistent pulmonary hypertension of the newborn. Am. J. Physiol. Lung Cell. Mol. Physiol. 2005, 289, L660–L666. [Google Scholar] [CrossRef] [Green Version]
  59. Jernigan, N.L.; Resta, T.C.; Walker, B.R. Contribution of oxygen radicals to altered NO-dependent pulmonary vasodilation in acute and chronic hypoxia. Am. J. Physiol. Lung Cell. Mol. Physiol. 2004, 286, L947–L955. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Nozik-Grayck, E.; Stenmark, K.R. Role of reactive oxygen species in chronic hypoxia-induced pulmonary hypertension and vascular remodeling. Adv. Exp. Med. Biol. 2007, 618, 101–112. [Google Scholar] [CrossRef] [PubMed]
  61. Ravi, Y.; Selvendiran, K.; Naidu, S.K.; Meduru, S.; Citro, L.A.; Bognár, B.; Khan, M.; Kálai, T.; Hideg, K.; Kuppusamy, M.L.; et al. Pulmonary hypertension secondary to left-heart failure involves peroxynitrite-induced downregulation of PTEN in the lung. Hypertension 2013, 61, 593–601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Zhao, Y.-Y.; Zhao, Y.D.; Mirza, M.K.; Huang, J.H.; Potula, H.-H.S.; Vogel, S.M.; Brovkovych, V.; Yuan, J.X.-J.; Wharton, J.; Malik, A.B. Persistent eNOS activation secondary to caveolin-1 deficiency induces pulmonary hypertension in mice and humans through PKG nitration. J. Clin. Investig. 2009, 119, 2009–2018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Oishi, P.; Grobe, A.; Benavidez, E.; Ovadia, B.; Harmon, C.; Ross, G.A.; Hendricks-Munoz, K.; Xu, J.; Black, S.M.; Fineman, J.R. Inhaled nitric oxide induced NOS inhibition and rebound pulmonary hypertension: A role for superoxide and peroxynitrite in the intact lamb. Am. J. Physiol. Lung Cell. Mol. Physiol. 2006, 290, L359–L366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Vara, D.; Pula, G. Reactive oxygen species: Physiological roles in the regulation of vascular cells. Curr. Mol. Med. 2014, 14, 1103–1125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Perez-Vizcaino, F.; Cogolludo, A.; Moreno, L. Reactive oxygen species signaling in pulmonary vascular smooth muscle. Respir. Physiol. Neurobiol. 2010, 174, 212–220. [Google Scholar] [CrossRef] [PubMed]
  66. Raaz, U.; Toh, R.; Maegdefessel, L.; Adam, M.; Nakagami, F.; Emrich, F.C.; Spin, J.M.; Tsao, P.S. Hemodynamic Regulation of Reactive Oxygen Species: Implications for Vascular Diseases. Antioxid. Redox Signal. 2014, 20, 914–928. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Krylatov, A.V.; Maslov, L.N.; Voronkov, N.; Boshchenko, A.; Popov, S.V.; Gomez, L.; Wang, H.; Jaggi, A.S.; Downey, J.M. Reactive Oxygen Species as Intracellular Signaling Molecules in the Cardiovascular System. Curr. Cardiol. Rev. 2018, 14, 290–300. [Google Scholar] [CrossRef]
  68. Burgoyne, J.R.; Mongue-Din, H.; Eaton, P.; Shah, A.M. Redox Signaling in Cardiac Physiology and Pathology. Circ. Res. 2012, 111, 1091–1106. [Google Scholar] [CrossRef]
  69. Burgoyne, J.R.; Oka, S.-I.; Ale-Agha, N.; Eaton, P. Hydrogen Peroxide Sensing and Signaling by Protein Kinases in the Cardiovascular System. Antioxid. Redox Signal. 2013, 18, 1042–1052. [Google Scholar] [CrossRef] [Green Version]
  70. Spickett, C.M.; Pitt, A.R.; Morrice, N.; Kolch, W. Proteomic analysis of phosphorylation, oxidation and nitrosylation in signal transduction. Biochim. Biophys. Acta 2006, 1764, 1823–1841. [Google Scholar] [CrossRef]
  71. Tang, H.; Viola, H.M.; Filipovska, A.; Hool, L.C. Cav1.2 calcium channel is glutathionylated during oxidative stress in guinea pig and ischemic human heart. Free Radic. Biol. Med. 2011, 51, 1501–1511. [Google Scholar] [CrossRef] [PubMed]
  72. Johnstone, V.P.A.; Hool, L.C. Glutathionylation of the L-type Ca2+ Channel in Oxidative Stress-Induced Pathology of the Heart. Int. J. Mol. Sci. 2014, 15, 19203–19225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Hawkins, B.J.; Irrinki, K.M.; Mallilankaraman, K.; Lien, Y.-C.; Wang, Y.; Bhanumathy, C.D.; Subbiah, R.; Ritchie, M.F.; Soboloff, J.; Baba, Y.; et al. S-glutathionylation activates STIM1 and alters mitochondrial homeostasis. J. Cell Biol. 2010, 190, 391–405. [Google Scholar] [CrossRef] [Green Version]
  74. Zha, X.-M.; Wang, R.; Collier, D.M.; Snyder, P.M.; Wemmie, J.A.; Welsh, M.J. Oxidant regulated inter-subunit disulfide bond formation between ASIC1a subunits. Proc. Natl. Acad. Sci. USA 2009, 106, 3573–3578. [Google Scholar] [CrossRef] [Green Version]
  75. Wu, X.; Hernandez-Enriquez, B.; Banas, M.; Xu, R.; Sesti, F. Molecular Mechanisms Underlying the Apoptotic Effect of KCNB1 K+ Channel Oxidation. J. Biol. Chem. 2013, 288, 4128–4134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Cotella, D.; Hernandez-Enriquez, B.; Wu, X.; Li, R.; Pan, Z.; Leveille, J.; Link, C.D.; Oddo, S.; Sesti, F. Toxic Role of K+ Channel Oxidation in Mammalian Brain. J. Neurosci. 2012, 32, 4133–4144. [Google Scholar] [CrossRef] [PubMed]
  77. Heo, J.; Raines, K.W.; Mocanu, V.; Campbell, S.L. Redox Regulation of RhoA. Biochemistry 2006, 45, 14481–14489. [Google Scholar] [CrossRef]
  78. Svoboda, L.K.; Reddie, K.G.; Zhang, L.; Vesely, E.D.; Williams, E.S.; Schumacher, S.M.; O’Connell, R.P.; Shaw, R.M.; Day, S.M.; Anumonwo, J.M.; et al. Redox-sensitive sulfenic acid modification regulates surface expression of the cardiovascular voltage-gated potassium channel Kv1.5. Circ. Res. 2012, 111, 842–853. [Google Scholar] [CrossRef] [Green Version]
  79. Panday, A.; Sahoo, M.K.; Osorio, D.; Batra, S. NADPH oxidases: An overview from structure to innate immunity-associated pathologies. Cell. Mol. Immunol. 2015, 12, 5–23. [Google Scholar] [CrossRef] [Green Version]
  80. Nguyen, G.T.; Green, E.R.; Mecsas, J. Neutrophils to the ROScue: Mechanisms of NADPH Oxidase Activation and Bacterial Resistance. Front. Cell Infect. Microbiol. 2017, 7, 373. [Google Scholar] [CrossRef] [PubMed]
  81. Menden, H.; Tate, E.; Hogg, N.; Sampath, V. LPS-mediated endothelial activation in pulmonary endothelial cells: Role of Nox2-dependent IKK-β phosphorylation. Am. J. Physiol. Lung Cell. Mol. Physiol. 2013, 304, L445–L455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Pendyala, S.; A Gorshkova, I.; Usatyuk, P.V.; He, D.; Pennathur, A.; Lambeth, J.D.; Thannickal, V.J.; Natarajan, V. Role of Nox4 and Nox2 in Hyperoxia-Induced Reactive Oxygen Species Generation and Migration of Human Lung Endothelial Cells. Antioxid. Redox Signal. 2009, 11, 747–764. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Ghouleh, I.A.; Sahoo, S.; Meijles, D.N.; Amaral, J.H.; De Jesus, D.S.; Sembrat, J.; Rojas, M.; Goncharov, D.A.; Goncharova, E.A.; Pagano, P.J. Endothelial Nox1 oxidase assembly in human pulmonary arterial hypertension; driver of Gremlin1-mediated proliferation. Clin. Sci. (Lond.) 2017, 131, 2019–2035. [Google Scholar] [CrossRef] [PubMed]
  84. Chignalia, A.Z.; Weinberg, G.; Dull, R. Norepinephrine Induces Lung Microvascular Endothelial Cell Death by NADPH Oxidase-Dependent Activation of Caspase-3. Oxidative Med. Cell. Longev. 2020, 2020, 2563764. [Google Scholar] [CrossRef]
  85. Hood, K.Y.; Mair, K.M.; Harvey, A.P.; Montezano, A.C.; Touyz, R.M.; MacLean, M.R. Serotonin Signaling Through the 5-HT(1B) Receptor and NADPH Oxidase 1 in Pulmonary Arterial Hypertension. Arterioscler. Thromb. Vasc. Biol. 2017, 37, 1361–1370. [Google Scholar] [CrossRef] [Green Version]
  86. Yu, W.; Ji, W.; Mi, L.; Lin, C. Mechanisms of N-acetylcysteine in reducing monocrotaline-induced pulmonary hypertension in rats: Inhibiting the expression of Nox1 in pulmonary vascular smooth muscle cells. Mol. Med. Rep. 2017, 16, 6148–6155. [Google Scholar] [CrossRef] [Green Version]
  87. Veit, F.; Pak, O.; Egemnazarov, B.; Roth, M.; Kosanovic, D.; Seimetz, M.; Sommer, N.; Ghofrani, H.-A.; Seeger, W.; Grimminger, F.; et al. Function of NADPH Oxidase 1 in Pulmonary Arterial Smooth Muscle Cells After Monocrotaline-Induced Pulmonary Vascular Remodeling. Antioxid. Redox Signal. 2013, 19, 2213–2231. [Google Scholar] [CrossRef]
  88. Iwata, K.; Ikami, K.; Matsuno, K.; Yamashita, T.; Shiba, D.; Ibi, M.; Misaki, M.; Masato, K.; Wenhao, C.; Jia, Z.; et al. Deficiency of NOX1/Nicotinamide Adenine Dinucleotide Phosphate, Reduced Form Oxidase Leads to Pulmonary Vascular Remodeling. Arterioscler. Thromb. Vasc. Biol. 2014, 34, 110–119. [Google Scholar] [CrossRef] [Green Version]
  89. Hendricks, K.; To, E.; Vlahos, R.; Broughton, B.; Peshavariya, H.; Selemidis, S. Influenza A virus causes vascular endothelial cell oxidative stress via NOX2 oxidase. Eur. Respir. J. 2016, 48 (Suppl. S60), PA3967. [Google Scholar] [CrossRef]
  90. Mittal, M.; Roth, M.; König, P.; Hofmann, S.; Dony, E.; Goyal, P.; Selbitz, A.-C.; Schermuly, R.T.; Ghofrani, H.A.; Kwapiszewska, G.; et al. Hypoxia-Dependent Regulation of Nonphagocytic NADPH Oxidase Subunit NOX4 in the Pulmonary Vasculature. Circ. Res. 2007, 101, 258–267. [Google Scholar] [CrossRef]
  91. Gandhirajan, R.K.; Meng, S.; Chandramoorthy, H.C.; Mallilankaraman, K.; Mancarella, S.; Gao, H.; Razmpour, R.; Yang, X.-F.; Houser, S.R.; Chen, J.; et al. Blockade of NOX2 and STIM1 signaling limits lipopolysaccharide-induced vascular inflammation. J. Clin. Investig. 2013, 123, 887–902. [Google Scholar] [CrossRef] [PubMed]
  92. Li, T.; Luo, X.-J.; Wang, E.-L.; Li, N.-S.; Zhang, X.-J.; Song, F.-L.; Yang, J.-F.; Liu, B.; Peng, J. Magnesium lithospermate B prevents phenotypic transformation of pulmonary arteries in rats with hypoxic pulmonary hypertension through suppression of NADPH oxidase. Eur. J. Pharmacol. 2019, 847, 32–41. [Google Scholar] [CrossRef] [PubMed]
  93. Bánfi, B.; Malgrange, B.; Knisz, J.; Steger, K.; Dubois-Dauphin, M.; Krause, K.-H. NOX3, a Superoxide-generating NADPH Oxidase of the Inner Ear. J. Biol. Chem. 2004, 279, 46065–46072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Goitre, L.; Distefano, P.V.; Moglia, A.; Nobiletti, N.; Baldini, E.; Trabalzini, L.; Keubel, J.; Trapani, E.; Shuvaev, V.V.; Muzykantov, V.R.; et al. Up-regulation of NADPH oxidase-mediated redox signaling contributes to the loss of barrier function in KRIT1 deficient endothelium. Sci. Rep. 2017, 7, 8296. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Schröder, K.; Zhang, M.; Benkhoff, S.; Mieth, A.; Pliquett, R.; Kosowski, J.; Kruse, C.; Luedike, P.; Michaelis, U.R.; Weissmann, N.; et al. Nox4 Is a Protective Reactive Oxygen Species Generating Vascular NADPH Oxidase. Circ. Res. 2012, 110, 1217–1225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. 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-β1 and insulin-like growth factor binding protein-3. Am. J. Physchol. Lung Cell. Mol. Physchol. 2009, 296, L489–L499. [Google Scholar] [CrossRef] [PubMed]
  97. Diebold, I.; Petry, A.; Hess, J.; Görlach, A. The NADPH Oxidase Subunit NOX4 Is a New Target Gene of the Hypoxia-inducible Factor-1. Mol. Biol. Cell 2010, 21, 2087–2096. [Google Scholar] [CrossRef] [Green Version]
  98. Hood, K.Y.; Montezano, A.C.; Harvey, A.P.; Nilsen, M.; MacLean, M.R.; Touyz, R.M. Nicotinamide Adenine Dinucleotide Phosphate Oxidase-Mediated Redox Signaling and Vascular Remodeling by 16α-Hydroxyestrone in Human Pulmonary Artery Cells: Implications in Pulmonary Arterial Hypertension. Hypertension 2016, 68, 796–808. [Google Scholar] [CrossRef]
  99. Martyn, K.D.; Frederick, L.M.; Von Loehneysen, K.; Dinauer, M.C.; Knaus, U.G. Functional analysis of Nox4 reveals unique characteristics compared to other NADPH oxidases. Cell. Signal. 2006, 18, 69–82. [Google Scholar] [CrossRef]
  100. Nisimoto, Y.; Jackson, H.M.; Ogawa, H.; Kawahara, T.; Lambeth, J.D. Constitutive NADPH-Dependent Electron Transferase Activity of the Nox4 Dehydrogenase Domain. Biochemistry 2010, 49, 2433–2442. [Google Scholar] [CrossRef]
  101. Nisimoto, Y.; Diebold, B.A.; Cosentino-Gomes, D.; Constentino-Gomes, D.; Lambeth, J.D. Nox4: A Hydrogen Peroxide-Generating Oxygen Sensor. Biochemistry 2014, 53, 5111–5120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Guo, X.; Fan, Y.; Cui, J.; Hao, B.; Zhu, L.; Sun, X.; He, J.; Yang, J.; Dong, J.; Wang, Y.; et al. NOX4 expression and distal arteriolar remodeling correlate with pulmonary hypertension in COPD. BMC Pulm. Med. 2018, 18, 111. [Google Scholar] [CrossRef]
  103. Veith, C.; Kraut, S.; Wilhelm, J.; Sommer, N.; Quanz, K.; Seeger, W.; Brandes, R.P.; Weissmann, N.; Schröder, K. NADPH oxidase 4 is not involved in hypoxia-induced pulmonary hypertension. Pulm. Circ. 2016, 6, 397–400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Ray, R.; Murdoch, C.E.; Wang, M.; Santos, C.X.; Zhang, M.; Alom-Ruiz, S.; Anilkumar, N.; Ouattara, A.; Cave, A.C.; Walker, S.J.; et al. Endothelial Nox4 NADPH Oxidase Enhances Vasodilatation and Reduces Blood Pressure In Vivo. Arter. Thromb. Vasc. Biol. 2011, 31, 1368–1376. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Zhang, M.; Mongue-Din, H.; Martin, D.; Catibog, N.; Smyrnias, I.; Zhang, X.; Yu, B.; Wang, M.; Brandes, R.P.; Schröder, K.; et al. Both cardiomyocyte and endothelial cell Nox4 mediate protection against hemodynamic overload-induced remodelling. Cardiovasc. Res. 2018, 114, 401–408. [Google Scholar] [CrossRef]
  106. Cadenas, E.; Davies, K.J.A. Mitochondrial free radical generation, oxidative stress, and aging. Free Radic. Biol. Med. 2000, 29, 222–230. [Google Scholar] [CrossRef]
  107. Miriyala, S.; Holley, A.K.; Clair, D.K.S. Mitochondrial superoxide dismutase--signals of distinction. Anti Cancer Agents Med. Chem. 2011, 11, 181–190. [Google Scholar] [CrossRef] [Green Version]
  108. Al Shahrani, M.; Heales, S.J.; Hargreaves, I.; Orford, M. Oxidative Stress: Mechanistic Insights into Inherited Mitochondrial Disorders and Parkinson’s Disease. J. Clin. Med. 2017, 6, 100. [Google Scholar] [CrossRef]
  109. Handy, D.E.; Lubos, E.; Yang, Y.; Galbraith, J.D.; Kelly, N.; Zhang, Y.-Y.; Leopold, J.A.; Loscalzo, J. Glutathione Peroxidase-1 Regulates Mitochondrial Function to Modulate Redox-dependent Cellular Responses. J. Biol. Chem. 2009, 284, 11913–11921. [Google Scholar] [CrossRef] [Green Version]
  110. Yu, E.; Mercer, J.; Bennett, M.R. Mitochondria in vascular disease. Cardiovasc. Res. 2012, 95, 173–182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Bai, J.; Cederbaum, A.I. Mitochondrial Catalase and Oxidative Injury. Biol. Signals Recept. 2001, 10, 189–199. [Google Scholar] [CrossRef] [PubMed]
  112. Murphy, M.P. How mitochondria produce reactive oxygen species. Biochem. J. 2009, 417, 1–13. [Google Scholar] [CrossRef] [Green Version]
  113. Suresh, K.; Servinsky, L.; Jiang, H.; Bigham, Z.; Yun, X.; Kliment, C.; Huetsch, J.; Damarla, M.; Shimoda, L.A. Reactive oxygen species induced Ca(2+) influx via TRPV4 and microvascular endothelial dysfunction in the SU5416/hypoxia model of pulmonary arterial hypertension. Am. J. Physiol. Lung Cell Mol. Physiol. 2018, 314, L893–L907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Sheak, J.R.; Yan, S.; Weise-Cross, L.; Ahmadian, R.; Walker, B.R.; Jernigan, N.L.; Resta, T.C. PKCβ and reactive oxygen species mediate enhanced pulmonary vasoconstrictor reactivity following chronic hypoxia in neonatal rats. Am. J. Physiol. Heart Circ. Physiol. 2020, 318, H470–H483. [Google Scholar] [CrossRef] [PubMed]
  115. Coggins, M.P.; Bloch, K.D.; Zadelaar, S.; Kleemann, R.; Verschuren, L.; Weij, J.D.V.-V.D.; Van Der Hoorn, J.; Princen, H.M.; Kooistra, T. Nitric Oxide in the Pulmonary Vasculature. Arter. Thromb. Vasc. Biol. 2007, 27, 1877–1885. [Google Scholar] [CrossRef] [Green Version]
  116. Vásquez-Vivar, J.; Kalyanaraman, B.; Martásek, P.; Hogg, N.; Masters, B.S.S.; Karoui, H.; Tordo, P.; Pritchard, K.A. Superoxide generation by endothelial nitric oxide synthase: The influence of cofactors. Proc. Natl. Acad. Sci. USA 1998, 95, 9220–9225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Schmidt, T.S.; Alp, N.J. Mechanisms for the role of tetrahydrobiopterin in endothelial function and vascular disease. Clin. Sci. (Lond.) 2007, 113, 47–63. [Google Scholar] [CrossRef] [Green Version]
  118. Griffith, O.W.; Stuehr, D.J. Nitric oxide synthases: Properties and catalytic mechanism. Annu. Rev. Physiol. 1995, 57, 707–736. [Google Scholar] [CrossRef] [PubMed]
  119. Habib, S.; Ali, A. Biochemistry of Nitric Oxide. Indian J. Clin. Biochem. IJCB 2011, 26, 3–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Vásquez-Vivar, J.; Kalyanaraman, B.; Martasek, P. The role of tetrahydrobiopterin in superoxide generation from eNOS: Enzymology and physiological implications. Free Radic. Res. 2003, 37, 121–127. [Google Scholar] [CrossRef] [PubMed]
  121. Crabtree, M.J.; Tatham, A.L.; Al-Wakeel, Y.; Warrick, N.; Hale, A.B.; Cai, S.; Channon, K.M.; Alp, N.J. Quantitative Regulation of Intracellular Endothelial Nitric-oxide Synthase (eNOS) Coupling by Both Tetrahydrobiopterin-eNOS Stoichiometry and Biopterin Redox Status: Insights From Cells With Tet-Regulated Gtp Cyclohydrolase I Expression. J. Biol. Chem. 2009, 284, 1136–1144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Crabtree, M.J.; Channon, K.M. Synthesis and recycling of tetrahydrobiopterin in endothelial function and vascular disease. Nitric Oxide Biol. Chem. 2011, 25, 81–88. [Google Scholar] [CrossRef] [Green Version]
  123. Vásquez-Vivar, J.; Martásek, P.; Whitsett, J.; Joseph, J.; Kalyanaraman, B. The ratio between tetrahydrobiopterin and oxidized tetrahydrobiopterin analogues controls superoxide release from endothelial nitric oxide synthase: An EPR spin trapping study. Biochem. J. 2002, 362 Pt 3, 733–739. [Google Scholar] [CrossRef]
  124. Khoo, J.P.; Zhao, L.; Alp, N.J.; Bendall, J.K.; Nicoli, T.; Rockett, K.; Wilkins, M.R.; Channon, K.M.; Woo, A.; Williams, W.G.; et al. Pivotal Role for Endothelial Tetrahydrobiopterin in Pulmonary Hypertension. Circulation 2005, 111, 2126–2133. [Google Scholar] [CrossRef] [Green Version]
  125. Bevers, L.M.; Braam, B.; Post, J.A.; Van Zonneveld, A.J.; Rabelink, T.J.; Koomans, H.A.; Verhaar, M.; Joles, J.A. Tetrahydrobiopterin, but Not l -Arginine, Decreases NO Synthase Uncoupling in Cells Expressing High Levels of Endothelial NO Synthase. Hypertension 2006, 47, 87–94. [Google Scholar] [CrossRef] [Green Version]
  126. Edgar, K.S.; Galvin, O.M.; Collins, A.; Katusic, Z.S.; McDonald, D.M. BH4-Mediated Enhancement of Endothelial Nitric Oxide Synthase Activity Reduces Hyperoxia-Induced Endothelial Damage and Preserves Vascular Integrity in the Neonate. Investig. Opthalmology Vis. Sci. 2017, 58, 230–241. [Google Scholar] [CrossRef] [Green Version]
  127. Schreiber, C.; Eilenberg, M.S.; Panzenboeck, A.; Winter, M.P.; Bergmeister, H.; Herzog, R.; Mascherbauer, J.; Lang, I.M.; Bonderman, D. Combined oral administration of L-arginine and tetrahydrobiopterin in a rat model of pulmonary arterial hypertension. Pulm. Circ. 2017, 7, 89–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Sheehy, A.M.; Burson, M.A.; Black, S.M. Nitric oxide exposure inhibits endothelial NOS activity but not gene expression: A role for superoxide. Am. J. Physiol. Lung Cell. Mol. Physiol. 1998, 274, L833–L841. [Google Scholar] [CrossRef] [PubMed]
  129. Wedgwood, S.; McMullan, D.M.; Bekker, J.M.; Fineman, J.R.; Black, S.M. Role for endothelin-1-induced superoxide and peroxynitrite production in rebound pulmonary hypertension associated with inhaled nitric oxide therapy. Circ. Res. 2001, 89, 357–364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Miller, O.; Tang, S.; Keech, A.; Celermajer, D. Rebound pulmonary hypertension on withdrawal from inhaled nitric oxide. Lancet 1995, 346, 51–52. [Google Scholar] [CrossRef]
  131. Atz, A.M.; Adatia, I.; Wessel, D.L. Rebound Pulmonary Hypertension After Inhalation of Nitric Oxide. Ann. Thorac. Surg. 1996, 62, 1759–1764. [Google Scholar] [CrossRef]
  132. Kostić, D.A.; Dimitrijević, D.S.; Stojanović, G.S.; Palić, I.R.; Đorđević, A.; Ickovski, J. Xanthine Oxidase: Isolation, Assays of Activity, and Inhibition. J. Chem. 2015, 2015, 294858. [Google Scholar] [CrossRef] [Green Version]
  133. Cantu-Medellin, N.; Kelley, E.E. Xanthine oxidoreductase-catalyzed reactive species generation: A process in critical need of reevaluation. Redox Biol. 2013, 1, 353–358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Kuwabara, Y.; Nishino, T.; Okamoto, K.; Matsumura, T.; Eger, B.T.; Pai, E.F.; Nishino, T. Unique amino acids cluster for switching from the dehydrogenase to oxidase form of xanthine oxidoreductase. Proc. Natl. Acad. Sci. USA 2003, 100, 8170–8175. [Google Scholar] [CrossRef] [Green Version]
  135. Battelli, M.G.; Polito, L.; Bortolotti, M.; Bolognesi, A. Xanthine Oxidoreductase in Drug Metabolism: Beyond a Role as a Detoxifying Enzyme. Curr. Med. Chem. 2016, 23, 4027–4036. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Chen, C.; Lü, J.-M.; Yao, Q. Hyperuricemia-Related Diseases and Xanthine Oxidoreductase (XOR) Inhibitors: An Overview. Med. Sci. Monit. 2016, 22, 2501–2512. [Google Scholar] [CrossRef] [Green Version]
  137. Kelley, E.E.; Khoo, N.K.; Hundley, N.J.; Malik, U.Z.; Freeman, B.A.; Tarpey, M.M. Hydrogen peroxide is the major oxidant product of xanthine oxidase. Free Radic. Biol. Med. 2010, 48, 493–498. [Google Scholar] [CrossRef] [Green Version]
  138. Terada, L.S.; Guidot, D.M.; Leff, J.A.; Willingham, I.R.; Hanley, M.E.; Piermattei, D.; Repine, J.E. Hypoxia injures endothelial cells by increasing endogenous xanthine oxidase activity. Proc. Natl. Acad. Sci. USA 1992, 89, 3362–3366. [Google Scholar] [CrossRef] [Green Version]
  139. Cheong, P.L.S. Effects of Xanthine Oxidase Inhibitors in Pulmonary Hypertension Associated with Chronic Lung Disease. Ph.D. Thesis, University of Dundee, Dundee, UK, 2019. [Google Scholar]
  140. Thorpe, L.W.; Westlund, K.N.; Kochersperger, L.M.; Abell, C.W.; Denney, R.M. Immunocytochemical localization of monoamine oxidases A and B in human peripheral tissues and brain. J. Histochem. Cytochem. 1987, 35, 23–32. [Google Scholar] [CrossRef] [Green Version]
  141. Shih, J.C.; Chen, K.; Ridd, M.J. Monoamine Oxidase: From Genes to Behavior. Annu. Rev. Neurosci. 1999, 22, 197–217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Fowler, C.J.; Mantle, T.J.; Tipton, K.F. The nature of the inhibition of rat liver monoamine oxidase types A and B by the acetylenic inhibitors clorgyline, l-deprenyl and pargyline. Biochem. Pharmacol. 1982, 31, 3555–3561. [Google Scholar] [CrossRef]
  143. Green, A.R.; Youdim, M.B. Effects of Monoamine Oxidase Inhibition by Clorgyline, Deprenil or Tranylcypromine on 5-Hydroxytryptamine Concentrations in Rat Brain and Hyperactivity Following Subsequent Tryptophan Administration. Br. J. Pharmacol. 1975, 55, 415–422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Fowler, C.J.; Benedetti, M.S. The Metabolism of Dopamine by Both Forms of Monoamine Oxidase in the Rat Brain and Its Inhibition by Cimoxatone. J. Neurochem. 1983, 40, 1534–1541. [Google Scholar] [CrossRef] [PubMed]
  145. O’Carroll, A.-M.; Fowler, C.J.; Phillips, J.P.; Tobbia, I.; Tipton, K.F. The deamination of dopamine by human brain monoamine oxidase. Specificity for the two enzyme forms in seven brain regions. Naunyn Schmiedeberg’s Arch. Pharmacol. 1983, 322, 198–202. [Google Scholar] [CrossRef] [PubMed]
  146. Tsugeno, Y.; Ito, A. A Key Amino Acid Responsible for Substrate Selectivity of Monoamine Oxidase A and B. J. Biol. Chem. 1997, 272, 14033–14036. [Google Scholar] [CrossRef] [Green Version]
  147. Gaweska, H.; Fitzpatrick, P.F. Structures and mechanism of the monoamine oxidase family. Biomol. Concepts 2011, 2, 365–377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Peña-Silva, R.A.; Miller, J.D.; Chu, Y.; Heistad, D.D. Serotonin produces monoamine oxidase-dependent oxidative stress in human heart valves. Am. J. Physiol. Heart Circ. Physiol. 2009, 297, H1354–H1360. [Google Scholar] [CrossRef] [Green Version]
  149. Zang, L.Y.; Misra, H.P. Generation of reactive oxygen species during the monoamine oxidase-catalyzed oxidation of the neurotoxicant, 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine. J. Biol. Chem. 1993, 268, 16504–16512. [Google Scholar]
  150. Sandri, G.; Panfili, E.; Ernster, L. Hydrogen peroxide production by monoamine oxidase in isolated rat-brain mitochondria: Its effect on glutathione levels and Ca2+ efflux. Biochim. Biophys. Acta 1990, 1035, 300–305. [Google Scholar] [CrossRef]
  151. Maurel-Ribes, A.; Hernandez, C.; Kunduzova, O.; Bompart, G.; Cambon, C.; Parini, A.; Francès, B. Age-dependent increase in hydrogen peroxide production by cardiac monoamine oxidase A in rats. Am. J. Physiol. Heart Circ. Physiol. 2003, 284, H1460–H1467. [Google Scholar] [CrossRef]
  152. Kunduzova, O.R.; Bianchi, P.; Parini, A.; Cambon, C. Hydrogen peroxide production by monoamine oxidase during ischemia/reperfusion. Eur. J. Pharmacol. 2002, 448, 225–230. [Google Scholar] [CrossRef]
  153. Simonson, S.G.; Zhang, J.; Canada, A.T.; Su, Y.-F.; Benveniste, H.; Piantadosi, C.A. Hydrogen Peroxide Production by Monoamine Oxidase during Ischemia-Reperfusion in the Rat Brain. J. Cereb. Blood Flow Metab. 1993, 13, 125–134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Wang, J.; Edmondson, D.E. Topological probes of monoamine oxidases A and B in rat liver mitochondria: Inhibition by TEMPO-substituted pargyline analogues and inactivation by proteolysis. Biochemistry 2011, 50, 2499–2505. [Google Scholar] [CrossRef] [Green Version]
  155. Mitoma, J.-Y.; Ito, A. Mitochondrial Targeting Signal of Rat Liver Monoamine Oxidase B Is Located at Its Carboxy Terminus. J. Biochem. 1992, 111, 20–24. [Google Scholar] [CrossRef] [Green Version]
  156. Son, S.-Y.; Ma, J.; Kondou, Y.; Yoshimura, M.; Yamashita, E.; Tsukihara, T. Structure of human monoamine oxidase A at 2.2-A resolution: The control of opening the entry for substrates/inhibitors. Proc. Natl. Acad. Sci. USA 2008, 105, 5739–5744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Sorato, E.; Menazza, S.; Zulian, A.; Sabatelli, P.; Gualandi, F.; Merlini, L.; Bonaldo, P.; Canton, M.; Bernardi, P.; Di Lisa, F. Monoamine oxidase inhibition prevents mitochondrial dysfunction and apoptosis in myoblasts from patients with collagen VI myopathies. Free Radic. Biol. Med. 2014, 75, 40–47. [Google Scholar] [CrossRef] [PubMed]
  158. Kaludercic, N.; Carpi, A.; Nagayama, T.; Sivakumaran, V.; Zhu, G.; Lai, E.W.; Bedja, D.; De Mario, A.; Chen, K.; Gabrielson, K.L.; et al. Monoamine Oxidase B Prompts Mitochondrial and Cardiac Dysfunction in Pressure Overloaded Hearts. Antioxid. Redox Signal 2014, 20, 267–280. [Google Scholar] [CrossRef] [Green Version]
  159. Antonucci, S.; Di Sante, M.; Tonolo, F.; Pontarollo, L.; Scalcon, V.; Alanova, P.; Menabo’, R.; Carpi, A.; Bindoli, A.; Rigobello, M.P.; et al. The Determining Role of Mitochondrial Reactive Oxygen Species Generation and Monoamine Oxidase Activity in Doxorubicin-Induced Cardiotoxicity. Antioxid. Redox Signal. 2020. [Google Scholar] [CrossRef]
  160. Jana, S.; Sinha, M.; Chanda, D.; Roy, T.; Banerjee, K.; Munshi, S.; Patro, B.S.; Chakrabarti, S. Mitochondrial dysfunction mediated by quinone oxidation products of dopamine: Implications in dopamine cytotoxicity and pathogenesis of Parkinson’s disease. Biochim. Biophys. Acta 2011, 1812, 663–673. [Google Scholar] [CrossRef] [Green Version]
  161. Sun, X.; Peters, E.; Schalij, I.; Andersen, S.; Bos, D.D.S.G.; Noordegraaf, A.V.; De Man, F.; Van Der Laarse, W.; Bogaard, H. Inhibition of Monoamine Oxidase-A Reduces Pulmonary Vascular Remodeling in Experimentally Induced Pulmonary Arterial Hypertension, in B108. In Under Pressure: The Role of Cellular Stress in Pulmonary Vascular Remodeling; American Thoracic Society: New York, NY, USA, 2019; p. A4205. [Google Scholar] [CrossRef]
  162. Sun, X.; Peters, E.; Schalij, I.; Andersen, S.; Bos, D.D.S.G.; Vonk-Noordegraaf, A.; De Man, F.; Van Der Laarse, W.; Bogaard, H. The effect of Monoamine oxidase A inhibition on experimentally induced pulmonary arterial hypertension. Eur. Respir. J. 2018, 52 (Suppl. S62), PA3072. [Google Scholar] [CrossRef]
  163. Archer, S.L.; Marsboom, G.; Kim, G.H.; Zhang, H.J.; Toth, P.T.; Svensson, E.C.; Dyck, J.R.; Gomberg-Maitland, M.; Thébaud, B.; Husain, A.N.; et al. Epigenetic attenuation of mitochondrial superoxide dismutase 2 in pulmonary arterial hypertension: A basis for excessive cell proliferation and a new therapeutic target. Circulation 2010, 121, 2661–2671. [Google Scholar] [CrossRef] [Green Version]
  164. Bonnet, S.; Michelakis, E.D.; Porter, C.J.; Andrade-Navarro, M.A.; Thébaud, B.; Bonnet, S.; Haromy, A.; Harry, G.; Moudgil, R.; McMurtry, M.S.; et al. An abnormal mitochondrial-hypoxia inducible factor-1alpha-Kv channel pathway disrupts oxygen sensing and triggers pulmonary arterial hypertension in fawn hooded rats: Similarities to human pulmonary arterial hypertension. Circulation 2006, 113, 2630–2641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Dennis, K.E.; Aschner, J.L.; Milatović, D.; Schmidt, J.W.; Aschner, M.; Kaplowitz, M.R.; Zhang, Y.; Fike, C.D. 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, L596–L607. [Google Scholar] [CrossRef] [PubMed]
  166. Ramiro-Diaz, J.M.; Nitta, C.H.; Maston, L.D.; Codianni, S.; Giermakowska, W.; Resta, T.C.; Bosc, L.V.G. NFAT is required for spontaneous pulmonary hypertension in superoxide dismutase 1 knockout mice. Am. J. Physiol. Lung Cell. Mol. Physiol. 2013, 304, L613–L625. [Google Scholar] [CrossRef] [Green Version]
  167. Brennan, L.A.; Steinhorn, R.; Wedgwood, S.; Mata-Greenwood, E.; Roark, E.A.; Russell, J.A.; Black, S.M. Increased Superoxide Generation Is Associated With Pulmonary Hypertension in Fetal Lambs. Circ. Res. 2003, 92, 683–691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Bowers, R.; Cool, C.; Murphy, R.C.; Tuder, R.M.; Hopken, M.W.; Flores, S.C.; Voelkel, N.F. Oxidative Stress in Severe Pulmonary Hypertension. Am. J. Respir. Crit. Care Med. 2004, 169, 764–769. [Google Scholar] [CrossRef] [PubMed]
  169. Alhawaj, R.; Patel, D.; Kelly, M.R.; Sun, N.; Wolin, M.S. Heme biosynthesis modulation via δ-aminolevulinic acid administration attenuates chronic hypoxia-induced pulmonary hypertension. Am. J. Physiol. Lung Cell. Mol. Physiol. 2015, 308, L719–L728. [Google Scholar] [CrossRef] [Green Version]
  170. Afolayan, A.J.; Eis, A.; Teng, R.-J.; Bakhutashvili, I.; Kaul, S.; Davis, J.M.; Konduri, G.G. Decreases in manganese superoxide dismutase expression and activity contribute to oxidative stress in persistent pulmonary hypertension of the newborn. Am. J. Physiol. Lung Cell. Mol. Physiol. 2012, 303, L870–L879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Zhang, Y.; Xu, J. miR-140-5p regulates hypoxia-mediated human pulmonary artery smooth muscle cell proliferation, apoptosis and differentiation by targeting Dnmt1 and promoting SOD2 expression. Biochem. Biophys. Res. Commun. 2016, 473, 342–348. [Google Scholar] [CrossRef]
  172. Tseng, V.; Ni, K.; Allawzi, A.; Prohaska, C.; Hernandez-Lagunas, L.; Elajaili, H.; Cali, V.; Midura, R.; Hascall, V.; Triggs-Raine, B.; et al. Extracellular Superoxide Dismutase Regulates Early Vascular Hyaluronan Remodeling in Hypoxic Pulmonary Hypertension. Sci. Rep. 2020, 10, 280. [Google Scholar] [CrossRef] [Green Version]
  173. Hartney, T.; Birari, R.; Venkataraman, S.; Villegas, L.; Martinez, M.; Black, S.M.; Stenmark, K.R.; Nozik-Gryck, E. Xanthine oxidase-derived ROS upregulate Egr-1 via ERK1/2 in PA smooth muscle cells; model to test impact of extracellular ROS in chronic hypoxia. PLoS ONE 2011, 6, e27531. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Nozik-Grayck, E.; Woods, C.; Taylor, J.M.; Benninger, R.K.P.; Johnson, R.D.; Villegas, L.R.; Stenmark, K.R.; Harrison, D.G.; Majka, S.M.; Irwin, D.; et al. Selective depletion of vascular EC-SOD augments chronic hypoxic pulmonary hypertension. Am. J. Physiol. Lung Cell. Mol. Physiol. 2014, 307, L868–L876. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Xu, D.; Guo, H.; Xu, X.; Chen, Y.; Fassett, J.; Hu, X.; Xu, Y.; Tang, Q.; Hu, D.; Somani, A.; et al. Exacerbated pulmonary arterial hypertension and right ventricular hypertrophy in animals with loss of function of extracellular superoxide dismutase. Hypertension 2011, 58, 303–309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Ducret, T.; El Arrouchi, J.; Courtois, A.; Quignard, J.-F.; Marthan, R.; Savineau, J.-P. Stretch-activated channels in pulmonary arterial smooth muscle cells from normoxic and chronically hypoxic rats. Cell Calcium 2010, 48, 251–259. [Google Scholar] [CrossRef]
  177. Wang, J.; Weigand, L.; Lu, W.; Sylvester, J.; Semenza, G.L.; Shimoda, L.A. Hypoxia Inducible Factor 1 Mediates Hypoxia-Induced TRPC Expression and Elevated Intracellular Ca 2+ in Pulmonary Arterial Smooth Muscle Cells. Circ. Res. 2006, 98, 1528–1537. [Google Scholar] [CrossRef] [Green Version]
  178. Lin, M.; Leung, G.P.; Zhang, W.-M.; Yang, X.-R.; Yip, K.-P.; Tse, C.-M.; Sham, J.S. Chronic hypoxia-induced upregulation of store-operated and receptor-operated Ca2+ channels in pulmonary arterial smooth muscle cells: A novel mechanism of hypoxic pulmonary hypertension. Circ. Res. 2004, 95, 496–505. [Google Scholar] [CrossRef] [Green Version]
  179. Chevalier, M.; Gilbert, G.; Roux, E.; Lory, P.; Marthan, R.; Savineau, J.-P.; Quignard, J.-F. T-type calcium channels are involved in hypoxic pulmonary hypertension. Cardiovasc. Res. 2014, 103, 597–606. [Google Scholar] [CrossRef] [Green Version]
  180. Shimoda, L.A.; Sham, J.S.K.; Shimoda, T.H.; Sylvester, J.T. L-type Ca(2+) channels, resting [Ca(2+)](i), and ET-1-induced responses in chronically hypoxic pulmonary myocytes. Am. J. Physiol. Lung Cell. Mol. Physiol. 2000, 279, L884–L894. [Google Scholar] [CrossRef]
  181. Nagaoka, T.; Morio, Y.; Casanova, N.; Bauer, N.; Gebb, S.; McMurtry, I.; Oka, M. Rho/Rho kinase signaling mediates increased basal pulmonary vascular tone in chronically hypoxic rats. Am. J. Physiol. Lung Cell. Mol. Physiol. 2004, 287, L665–L672. [Google Scholar] [CrossRef]
  182. Weigand, L.; Sylvester, J.T.; Shimoda, L.A. Mechanisms of endothelin-1-induced contraction in pulmonary arteries from chronically hypoxic rats. Am. J. Physiol. Lung Cell. Mol. Physiol. 2006, 290, L284–L290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Catterall, W.A. Voltage-gated calcium channels. Cold Spring Harb. Perspect. Biol. 2011, 3, a003947. [Google Scholar] [CrossRef] [PubMed]
  184. Firth, A.L.; Remillard, C.V.; Platoshyn, O.; Fantozzi, I.; Ko, E.A.; Yuan, J.X.-J. Functional Ion Channels in Human Pulmonary Artery Smooth Muscle Cells: Voltage-Dependent Cation Channels. Pulm. Circ. 2011, 1, 48–71. [Google Scholar] [CrossRef] [Green Version]
  185. Makino, A.; Firth, A.L.; Yuan, J.X.-J. Endothelial and smooth muscle cell ion channels in pulmonary vasoconstriction and vascular remodeling. Compr. Physiol. 2011, 1, 1555–1602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Shimoda, L.A.; Sylvester, J.T.; Sham, J.S.K. Chronic hypoxia alters effects of endothelin and angiotensin on K+ currents in pulmonary arterial myocytes. Am. J. Physiol. 1999, 277, L431–L439. [Google Scholar] [CrossRef]
  187. Reeve, H.L.; Michelakis, E.; Nelson, D.P.; Weir, E.K.; Archer, S.L. Alterations in a redox oxygen sensing mechanism in chronic hypoxia. J. Appl. Physiol. 2001, 90, 2249–2256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Yuan, J.X.-J.; Aldinger, A.M.; Juhaszova, M.; Wang, J.; Conte, J.V.; Gaine, S.P.; Orens, J.B.; Rubin, L.J. Dysfunctional Voltage-Gated K + Channels in Pulmonary Artery Smooth Muscle Cells of Patients With Primary Pulmonary Hypertension. Circulation 1998, 98, 1400–1406. [Google Scholar] [CrossRef] [Green Version]
  189. Lai, N.; Lu, W.; Wang, J. Ca2+ and ion channels in hypoxia-mediated pulmonary hypertension. Int. J. Clin. Exp. Pathol. 2015, 8, 1081–1092. [Google Scholar]
  190. Liou, J.; Kim, M.L.; Heo, W.D.; Jones, J.T.; Myers, J.W.; Ferrell, J.E.; Meyer, T. STIM Is a Ca2+ Sensor Essential for Ca2+-Store-Depletion-Triggered Ca2+ Influx. Curr. Biol. 2005, 15, 1235–1241. [Google Scholar] [CrossRef] [Green Version]
  191. Hewavitharana, T.; Deng, X.; Soboloff, J.; Gill, D.L. Role of STIM and Orai proteins in the store-operated calcium signaling pathway. Cell Calcium 2007, 42, 173–182. [Google Scholar] [CrossRef]
  192. Garcia, S.M.; Herbert, L.M.; Walker, B.R.; Resta, T.C.; Jernigan, N.L. Coupling of store-operated calcium entry to vasoconstriction is acid-sensing ion channel 1a dependent in pulmonary but not mesenteric arteries. PLoS ONE 2020, 15, e0236288. [Google Scholar] [CrossRef]
  193. Salido, G.M.; Jardin, I.; Rosado, J.A. The TRPC ion channels: Association with Orai1 and STIM1 proteins and participation in capacitative and non-capacitative calcium entry. Adv. Exp. Med. Biol. 2011, 704, 413–433. [Google Scholar] [CrossRef] [PubMed]
  194. Snow, J.B.; Kanagy, N.L.; Walker, B.R.; Resta, T.C. Rat Strain Differences in Pulmonary Artery Smooth Muscle Ca2+Entry Following Chronic Hypoxia. Microcirculation 2009, 16, 603–614. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Hirenallur-S, D.K.; Haworth, S.T.; Leming, J.T.; Chang, J.; Hernández, G.; Gordon, J.B.; Rusch, N.J. Upregulation of vascular calcium channels in neonatal piglets with hypoxia-induced pulmonary hypertension. Am. J. Physiol. Lung Cell. Mol. Physiol. 2008, 295, L915–L924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Wan, J.; Yamamura, A.; Zimnicka, A.M.; Voiriot, G.; Smith, K.A.; Tang, H.; Ayon, R.J.; Choudhury, M.S.R.; Ko, E.A.; Wang, J.; et al. Chronic hypoxia selectively enhances L- and T-type voltage-dependent Ca2+ channel activity in pulmonary artery by upregulating Cav1.2 and Cav3.2. Am. J. Physiol. Lung Cell. Mol. Physiol. 2013, 305, L154–L164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Oka, M.; Morris, K.G.; McMurtry, I.F. NIP-121 is more effective than nifedipine in acutely reversing chronic pulmonary hypertension. J. Appl. Physiol. 1993, 75, 1075–1080. [Google Scholar] [CrossRef] [PubMed]
  198. Johnson, D.C.; Joshi, R.C.; Mehta, R.; Cunnington, A.R. Acute and long term effect of nifedipine on pulmonary hypertension secondary to chronic obstructive airways disease. Eur. J. Respir. Dis. Suppl. 1986, 146, 495–502. [Google Scholar]
  199. Brown, S.E.; Linden, G.S.; King, R.R.; Blair, G.P.; Stansbury, D.W.; Light, R.W. Effects of verapamil on pulmonary haemodynamics during hypoxaemia, at rest, and during exercise in patients with chronic obstructive pulmonary disease. Thorax 1983, 38, 840–844. [Google Scholar] [CrossRef] [Green Version]
  200. Chen, T.-X.; Xu, X.-Y.; Zhao, Z.; Zhao, F.-Y.; Gao, Y.-M.; Yan, X.-H.; Wan, Y. Hydrogen peroxide is a critical regulator of the hypoxia-induced alterations of store-operated Ca2+ entry into rat pulmonary arterial smooth muscle cells. Am. J. Physiol. Lung Cell. Mol. Physiol. 2017, 312, L477–L487. [Google Scholar] [CrossRef]
  201. Malczyk, M.; Veith, C.; Fuchs, B.; Hofmann, K.; Storch, U.; Schermuly, R.T.; Witzenrath, M.; Ahlbrecht, K.; Fecher-Trost, C.; Flockerzi, V.; et al. Classical Transient Receptor Potential Channel 1 in Hypoxia-induced Pulmonary Hypertension. Am. J. Respir. Crit. Care Med. 2013, 188, 1451–1459. [Google Scholar] [CrossRef]
  202. Xia, Y.; Yang, X.R.; Fu, Z.; Paudel, O.; Abramowitz, J.; Birnbaumer, L.; Sham, J.S. Classical transient receptor potential 1 and 6 contribute to hypoxic pulmonary hypertension through differential regulation of pulmonary vascular functions. Hypertension 2014, 63, 173–180. [Google Scholar] [CrossRef] [Green Version]
  203. Smith, K.A.; Voiriot, G.; Tang, H.; Fraidenburg, D.R.; Song, S.; Yamamura, H.; Yamamura, A.; Guo, Q.; Wan, J.; Pohl, N.M.; et al. Notch Activation of Ca2+Signaling in the Development of Hypoxic Pulmonary Vasoconstriction and Pulmonary Hypertension. Am. J. Respir. Cell Mol. Biol. 2015, 53, 355–367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Yang, X.-R.; Lin, A.H.Y.; Hughes, J.M.; Flavahan, N.A.; Cao, Y.-N.; Liedtke, W.; Sham, J.S.K. Upregulation of osmo-mechanosensitive TRPV4 channel facilitates chronic hypoxia-induced myogenic tone and pulmonary hypertension. Am. J. Physiol. Lung Cell. Mol. Physiol. 2012, 302, L555–L568. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Dahan, D.; Ducret, T.; Quignard, J.-F.; Marthan, R.; Savineau, J.-P.; Estève, E. Implication of the ryanodine receptor in TRPV4-induced calcium response in pulmonary arterial smooth muscle cells from normoxic and chronically hypoxic rats. Am. J. Physiol. Lung Cell. Mol. Physiol. 2012, 303, L824–L833. [Google Scholar] [CrossRef] [Green Version]
  206. Xia, Y.; Fu, Z.; Hu, J.; Huang, C.; Paudel, O.; Cai, S.; Liedtke, W.; Sham, J.S.K. TRPV4 channel contributes to serotonin-induced pulmonary vasoconstriction and the enhanced vascular reactivity in chronic hypoxic pulmonary hypertension. Am. J. Physiol. Cell Physiol. 2013, 305, C704–C715. [Google Scholar] [CrossRef] [Green Version]
  207. Nitta, C.H.; Osmond, D.A.; Herbert, L.M.; Beasley, B.F.; Resta, T.C.; Walker, B.R.; Jernigan, N.L. Role of ASIC1 in the development of chronic hypoxia-induced pulmonary hypertension. Am. J. Physiol. Heart Circ. Physiol. 2014, 306, H41–H52. [Google Scholar] [CrossRef] [Green Version]
  208. He, X.; Song, S.; Ayon, R.J.; Balisterieri, A.; Black, S.M.; Makino, A.; Gil Wier, W.; Zang, W.-J.; Yuan, J.X.-J. Hypoxia selectively upregulates cation channels and increases cytosolic [Ca2+] in pulmonary, but not coronary, arterial smooth muscle cells. Am. J. Physiol. Cell Physiol. 2018, 314, C504–C517. [Google Scholar] [CrossRef] [Green Version]
  209. Song, M.Y.; Makino, A.; Yuan, J.X.-J. STIM2 Contributes to Enhanced Store-Operated Ca2+ Entry in Pulmonary Artery Smooth Muscle Cells from Patients with Idiopathic Pulmonary Arterial Hypertension. Pulm. Circ. 2011, 1, 84–94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Campbell, D.L.; Stamler, J.S.; Strauss, H.C. Redox modulation of L-type calcium channels in ferret ventricular myocytes. Dual mechanism regulation by nitric oxide and S-nitrosothiols. J. Gen. Physiol. 1996, 108, 277–293. [Google Scholar] [CrossRef] [PubMed]
  211. Hu, H.; Chiamvimonvat, N.; Yamagishi, T.; Marban, E. Direct inhibition of expressed cardiac L-type Ca2+ channels by S-nitrosothiol nitric oxide donors. Circ. Res. 1997, 81, 742–752. [Google Scholar] [CrossRef]
  212. Viola, H.M.; Arthur, P.G.; Hool, L.C. Transient Exposure to Hydrogen Peroxide Causes an Increase in Mitochondria-Derived Superoxide As a Result of Sustained Alteration in L-Type Ca2+ Channel Function in the Absence of Apoptosis in Ventricular Myocytes. Circ. Res. 2007, 100, 1036–1044. [Google Scholar] [CrossRef] [Green Version]
  213. Akaishi, T.; Nakazawa, K.; Sato, K.; Saito, H.; Ohno, Y.; Ito, Y. Hydrogen peroxide modulates whole cell Ca2+ currents through L-type channels in cultured rat dentate granule cells. Neurosci. Lett. 2004, 356, 25–28. [Google Scholar] [CrossRef] [PubMed]
  214. Yang, L.; Xu, J.; Minobe, E.; Yu, L.; Feng, R.; Kameyama, A.; Yazawa, K.; Kameyama, M. Mechanisms underlying the modulation of L-type Ca2+ channel by hydrogen peroxide in guinea pig ventricular myocytes. J. Physiol. Sci. 2013, 63, 419–426. [Google Scholar] [CrossRef]
  215. Luke, T.; Maylor, J.; Undem, C.; Sylvester, J.T.; Shimoda, L.A. Kinase-dependent activation of voltage-gated Ca2+ channels by ET-1 in pulmonary arterial myocytes during chronic hypoxia. Am. J. Physiol. Lung Cell. Mol. Physiol. 2012, 302, L1128–L1139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Cosentino-Gomes, D.; Rocco-Machado, N.; Meyer-Fernandes, J.R. Cell Signaling through Protein Kinase C Oxidation and Activation. Int. J. Mol. Sci. 2012, 13, 10697–10721. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Wu, W.; Platoshyn, O.; Firth, A.L.; Yuan, J.X.-J. Hypoxia divergently regulates production of reactive oxygen species in human pulmonary and coronary artery smooth muscle cells. Am. J. Physiol. Lung Cell. Mol. Physiol. 2007, 293, L952–L959. [Google Scholar] [CrossRef] [Green Version]
  218. Wedgwood, S.; Black, S.M. Endothelin-1 decreases endothelial NOS expression and activity through ETA receptor-mediated generation of hydrogen peroxide. Am. J. Physiol. Lung Cell. Mol. Physiol. 2005, 288, L480–L487. [Google Scholar] [CrossRef] [Green Version]
  219. Zeng, Q.; Zhou, Q.; Yao, F.; O’Rourke, S.T.; Sun, C. Endothelin-1 Regulates Cardiac L-Type Calcium Channels via NAD(P)H Oxidase-Derived Superoxide. J. Pharmacol. Exp. Ther. 2008, 326, 732–738. [Google Scholar] [CrossRef] [Green Version]
  220. Weissmann, N.; Dietrich, A.; Fuchs, B.; Kalwa, H.; Ay, M.; Dumitrascu, R.; Olschewski, A.; Storch, U.; Schnitzler, M.M.Y.; Ghofrani, H.-A.; et al. Classical transient receptor potential channel 6 (TRPC6) is essential for hypoxic pulmonary vasoconstriction and alveolar gas exchange. Proc. Natl. Acad. Sci. USA 2006, 103, 19093–19098. [Google Scholar] [CrossRef] [Green Version]
  221. Aires, V.; Hichami, A.; Boulay, G.; Khan, N.A. Activation of TRPC6 calcium channels by diacylglycerol (DAG)-containing arachidonic acid: A comparative study with DAG-containing docosahexaenoic acid. Biochimie 2007, 89, 926–937. [Google Scholar] [CrossRef]
  222. Weissmann, N.; Sydykov, A.; Kalwa, H.; Storch, U.; Fuchs, B.; Schnitzler, M.M.Y.; Brandes, R.P.; Grimminger, F.; Meissner, M.; Freichel, M.; et al. Activation of TRPC6 channels is essential for lung ischaemia–reperfusion induced oedema in mice. Nat. Commun. 2012, 3, 649. [Google Scholar] [CrossRef]
  223. Inoue, R.; Jensen, L.J.; Jian, Z.; Shi, J.; Hai, L.; Lurie, A.I.; Henriksen, F.H.; Salomonsson, M.; Morita, H.; Kawarabayashi, Y.; et al. Synergistic activation of vascular TRPC6 channel by receptor and mechanical stimulation via phospholipase C/diacylglycerol and phospholipase A2/omega-hydroxylase/20-HETE pathways. Circ. Res. 2009, 104, 1399–1409. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Ding, Y.; Winters, A.; Ding, M.; Graham, S.; Akopova, I.; Muallem, S.; Wang, Y.; Hong, J.H.; Gryczynski, Z.; Yang, S.-H.; et al. Reactive Oxygen Species-mediated TRPC6 Protein Activation in Vascular Myocytes, a Mechanism for Vasoconstrictor-regulated Vascular Tone. J. Biol. Chem. 2011, 286, 31799–31809. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Marziano, C.; Hong, K.; Cope, E.L.; Kotlikoff, M.I.; Isakson, B.E.; Sonkusare, S.K. Nitric Oxide–Dependent Feedback Loop Regulates Transient Receptor Potential Vanilloid 4 (TRPV4) Channel Cooperativity and Endothelial Function in Small Pulmonary Arteries. J. Am. Heart Assoc. 2017, 6, e007157. [Google Scholar] [CrossRef] [Green Version]
  226. Sukumaran, S.V.; Singh, T.U.; Parida, S.; Reddy, C.N.; Thangamalai, R.; Kandasamy, K.; Singh, V.; Mishra, S.K. TRPV4 channel activation leads to endothelium-dependent relaxation mediated by nitric oxide and endothelium-derived hyperpolarizing factor in rat pulmonary artery. Pharmacol. Res. 2013, 78, 18–27. [Google Scholar] [CrossRef]
  227. Ottolini, M.; Daneva, Z.; Chen, Y.L.; Cope, E.L.; Kasetti, R.B.; Zode, G.S.; Sonkusare, S.K. Mechanisms underlying selective coupling of endothelial Ca(2+) signals with eNOS vs. IK/SK channels in systemic and pulmonary arteries. J. Physiol. 2020, 598, 3577–3596. [Google Scholar] [CrossRef] [PubMed]
  228. Cussac, L.-A.; Cardouat, G.; Pillai, N.T.; Campagnac, M.; Robillard, P.; Montillaud, A.; Guibert, C.; Gailly, P.; Marthan, R.; Quignard, J.-F.; et al. TRPV4 channel mediates adventitial fibroblast activation and adventitial remodeling in pulmonary hypertension. Am. J. Physiol. Lung Cell. Mol. Physiol. 2020, 318, L135–L146. [Google Scholar] [CrossRef] [PubMed]
  229. Pankey, E.A.; Zsombok, A.; Lasker, G.F.; Kadowitz, P.J. Analysis of responses to the TRPV4 agonist GSK1016790A in the pulmonary vascular bed of the intact-chest rat. Am. J. Physiol. Heart Circ. Physiol. 2014, 306, 33–40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  230. Suresh, K.; Servinsky, L.; Reyes, J.; Baksh, S.; Undem, C.; Caterina, M.; Pearse, D.B.; Shimoda, L.A. Hydrogen peroxide-induced calcium influx in lung microvascular endothelial cells involves TRPV4. Am. J. Physiol. Lung Cell. Mol. Physiol. 2015, 309, L1467–L1477. [Google Scholar] [CrossRef] [Green Version]
  231. Suresh, K.; Servinsky, L.; Reyes, J.; Undem, C.; Zaldumbide, J.; Rentsendorj, O.; Modekurty, S.; Dodd-O, J.M.; Scott, A.L.; Pearse, D.B.; et al. CD36 mediates H2O2-induced calcium influx in lung microvascular endothelial cells. Am. J. Physiol. Lung Cell. Mol. Physiol. 2017, 312, L143–L153. [Google Scholar] [CrossRef] [Green Version]
  232. Cao, S.; Anishkin, A.; Zinkevich, N.S.; Nishijima, Y.; Korishettar, A.; Wang, Z.; Fang, J.; Wilcox, D.A.; Zhang, D.X. Transient receptor potential vanilloid 4 (TRPV4) activation by arachidonic acid requires protein kinase A–mediated phosphorylation. J. Biol. Chem. 2018, 293, 5307–5322. [Google Scholar] [CrossRef] [Green Version]
  233. Suresh, K.; Servinsky, L.; Jiang, H.; Bigham, Z.; Zaldumbide, J.; Huetsch, J.C.; Kliment, C.; Acoba, M.G.; Kirsch, B.J.; Claypool, S.M.; et al. Regulation of mitochondrial fragmentation in microvascular endothelial cells isolated from the SU5416/hypoxia model of pulmonary arterial hypertension. Am. J. Physiol. Lung. Cell Mol. Physiol. 2019, 317, L639–L652. [Google Scholar] [CrossRef] [PubMed]
  234. Jernigan, N.L. Smooth muscle acid-sensing ion channel 1: Pathophysiological implication in hypoxic pulmonary hypertension. Exp. Physiol. 2015, 100, 111–120. [Google Scholar] [CrossRef] [PubMed]
  235. Sherwood, T.W.; Frey, E.N.; Askwith, C.C. Structure and activity of the acid-sensing ion channels. Am. J. Physiol. Cell Physiol. 2012, 303, C699–C710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Mukhopadhyay, M.; Bera, A.K. Modulation of acid-sensing ion channels by hydrogen sulfide. Biochem. Biophys. Res. Commun. 2020, 527, 71–75. [Google Scholar] [CrossRef] [PubMed]
  237. Andrey, F.; Tsintsadze, T.; Volkova, T.; Lozovaya, N.; Krishtal, O. Acid sensing ionic channels: Modulation by redox reagents. Biochim. Biophys. Acta 2005, 1745, 1–6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Cho, J.-H.; Askwith, C.C. Potentiation of acid-sensing ion channels by sulfhydryl compounds. Am. J. Physiol. Cell Physiol. 2007, 292, C2161–C2174. [Google Scholar] [CrossRef] [PubMed]
  239. Chu, X.-P.; Close, N.; Saugstad, J.A.; Xiong, Z.-G. ASIC1a-specific modulation of acid-sensing ion channels in mouse cortical neurons by redox reagents. J. Neurosci. 2006, 26, 5329–5339. [Google Scholar] [CrossRef] [Green Version]
  240. Mungai, P.T.; Waypa, G.B.; Jairaman, A.; Prakriya, M.; Dokic, D.; Ball, M.K.; Schumacker, P.T. Hypoxia Triggers AMPK Activation through Reactive Oxygen Species-Mediated Activation of Calcium Release-Activated Calcium Channels. Mol. Cell. Biol. 2011, 31, 3531–3545. [Google Scholar] [CrossRef] [Green Version]
  241. Dyachenko, V.; Rueckschloss, U.; Isenberg, G. Modulation of cardiac mechanosensitive ion channels involves superoxide, nitric oxide and peroxynitrite. Cell Calcium 2009, 45, 55–64. [Google Scholar] [CrossRef]
  242. Burg, E.D.; Remillard, C.V.; Yuan, J.X.-J. Potassium channels in the regulation of pulmonary artery smooth muscle cell proliferation and apoptosis: Pharmacotherapeutic implications. Br. J. Pharmacol. 2008, 153 (Suppl. S1), S99–S111. [Google Scholar] [CrossRef] [Green Version]
  243. Weise-Cross, L.; Resta, T.C.; Jernigan, N.L. Redox Regulation of Ion Channels and Receptors in Pulmonary Hypertension. Antioxid. Redox Signal. 2019, 31, 898–915. [Google Scholar] [CrossRef] [PubMed]
  244. Ayon, R.J.; Tang, H.; Fernandez, R.A.; Makino, A.; Yuan, J.X.-J. Smooth Muscle Cell Ion Channels in Pulmonary Arterial Hypertension: Pathogenic Role in Pulmonary Vasoconstriction and Vascular Remodeling. In Vascular Ion Channels in Physiology and Disease; Levitan, P.I., Dopico, M.D.P.A.M., Eds.; Springer International Publishing: Cham, Switzerland, 2016; pp. 295–324. [Google Scholar] [CrossRef]
  245. Doi, S.; Damron, D.S.; Ogawa, K.; Tanaka, S.; Horibe, M.; A Murray, P. K(+) channel inhibition, calcium signaling, and vasomotor tone in canine pulmonary artery smooth muscle. Am. J. Physiol. Lung Cell. Mol. Physiol. 2000, 279, L242–L251. [Google Scholar] [CrossRef] [PubMed]
  246. Joshi, S.; Balan, P.; Gurney, A.M. Pulmonary vasoconstrictor action of KCNQ potassium channel blockers. Respir. Res. 2006, 7, 31. [Google Scholar] [CrossRef] [Green Version]
  247. Wang, J.; Juhaszova, M.; Rubin, L.J.; Yuan, X.J. Hypoxia inhibits gene expression of voltage-gated K+ channel alpha subunits in pulmonary artery smooth muscle cells. J. Clin. Investig. 1997, 100, 2347–2353. [Google Scholar] [CrossRef] [Green Version]
  248. Wang, J.; Weigand, L.; Wang, W.; Sylvester, J.T.; Shimoda, L.A. Chronic hypoxia inhibits Kv channel gene expression in rat distal pulmonary artery. Am. J. Physiol. Lung Cell. Mol. Physiol. 2005, 288, L1049–L1058. [Google Scholar] [CrossRef]
  249. Mittal, M.; Gu, X.Q.; Pak, O.; Pamenter, M.E.; Haag, D.; Fuchs, D.B.; Schermuly, R.T.; Ghofrani, H.-A.; Brandes, R.P.; Seeger, W.; et al. Hypoxia induces Kv channel current inhibition by increased NADPH oxidase-derived reactive oxygen species. Free Radic. Biol. Med. 2012, 52, 1033–1042. [Google Scholar] [CrossRef]
  250. Platoshyn, O.; Yu, Y.; Golovina, V.A.; McDaniel, S.S.; Krick, S.; Li, L.; Wang, J.-Y.; Rubin, L.J.; Yuan, J.X.-J. Chronic hypoxia decreases KV channel expression and function in pulmonary artery myocytes. Am. J. Physiol. Cell. Mol. Physiol. 2001, 280, L801–L812. [Google Scholar] [CrossRef] [Green Version]
  251. Fan, Z.; Liu, B.; Zhang, S.; Liu, H.; Li, Y.; Wang, N.; Liu, Y.; Li, J.; Wang, N.; Liu, Y.; et al. YM155, a selective survivin inhibitor, reverses chronic hypoxic pulmonary hypertension in rats via upregulating voltage-gated potassium channels. Clin. Exp. Hypertens. 2015, 37, 381–387. [Google Scholar] [CrossRef]
  252. Michelakis, E.D.; McMurtry, M.S.; Wu, X.-C.; Dyck, J.R.B.; Moudgil, R.; Hopkins, T.A.; Lopaschuk, G.D.; Puttagunta, L.; Waite, R.; Archer, S.L. Dichloroacetate, a metabolic modulator, prevents and reverses chronic hypoxic pulmonary hypertension in rats: Role of increased expression and activity of voltage-gated potassium channels. Circulation 2002, 105, 244–250. [Google Scholar] [CrossRef] [Green Version]
  253. Zheng, L.; Liu, M.; Wei, M.; Liu, Y.; Dong, M.; Luo, Y.; Zhao, P.; Dong, H.; Niu, W.; Yan, Z.; et al. Tanshinone IIA attenuates hypoxic pulmonary hypertension via modulating KV currents. Respir. Physiol. Neurobiol. 2015, 205, 120–128. [Google Scholar] [CrossRef]
  254. Frazzini, V.; Guarnieri, S.; Bomba, M.; Navarra, R.; Morabito, C.; A Mariggiò, M.; Sensi, S.L. Altered Kv2.1 functioning promotes increased excitability in hippocampal neurons of an Alzheimer’s disease mouse model. Cell Death Dis. 2016, 7, e2100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Morecroft, I.; Murray, A.; Nilsen, M.; Gurney, A.M.; MacLean, M.R. Treatment with the Kv7 potassium channel activator flupirtine is beneficial in two independent mouse models of pulmonary hypertension. Br. J. Pharmacol. 2009, 157, 1241–1249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Zuo, X.; Zong, F.; Wang, H.; Wang, Q.; Xie, W.; Wang, H. Iptakalim, a novel ATP-sensitive potassium channel opener, inhibits pulmonary arterial smooth muscle cell proliferation by downregulation of PKC-α. J. Biomed. Res. 2011, 25, 392–401. [Google Scholar] [CrossRef] [Green Version]
  257. Revermann, M.; Neofitidou, S.; Kirschning, T.; Schloss, M.; Brandes, R.P.; Hofstetter, C. Inhalation of the BK(Ca)-opener NS1619 attenuates right ventricular pressure and improves oxygenation in the rat monocrotaline model of pulmonary hypertension. PLoS ONE 2014, 9, e86636. [Google Scholar] [CrossRef] [PubMed]
  258. Roth, M.; Rupp, M.; Hofmann, S.; Mittal, M.; Fuchs, B.; Sommer, N.; Parajuli, N.; Quanz, K.; Schubert, D.; Dony, E.; et al. Heme oxygenase-2 and large-conductance Ca2+-activated K+ channels: Lung vascular effects of hypoxia. Am. J. Respir. Crit. Care Med. 2009, 180, 353–364. [Google Scholar] [CrossRef]
  259. Nielsen, G.; Wandall-Frostholm, C.; Sadda, V.; Oliván-Viguera, A.; Lloyd, E.E.; Bryan, R.M.; Simonsen, U.; Köhler, R. Alterations of N-3 polyunsaturated fatty acid-activated K2P channels in hypoxia-induced pulmonary hypertension. Basic Clin. Pharmacol. Toxicol. 2013, 113, 250–258. [Google Scholar] [CrossRef]
  260. Pandit, L.M.; Lloyd, E.E.; Reynolds, J.O.; Lawrence, W.S.; Reynolds, C.; Wehrens, X.H.; Bryan, R.M. TWIK-2 Channel Deficiency Leads to Pulmonary Hypertension Through a Rho-Kinase–Mediated Process. Hypertension 2014, 64, 1260–1265. [Google Scholar] [CrossRef] [Green Version]
  261. Kitagawa, M.G.; Reynolds, J.O.; Durgan, D.; Rodney, G.; Karmouty-Quintana, H.; Bryan, R.; Pandit, L.M. Twik-2(-/-) mouse demonstrates pulmonary vascular heterogeneity in intracellular pathways for vasocontractility. Physiol. Rep. 2019, 7, e13950. [Google Scholar] [CrossRef] [Green Version]
  262. Antigny, F.; Hautefort, A.; Meloche, J.; Belacel-Ouari, M.; Manoury, B.; Rucker-Martin, C.; Péchoux, C.; Potus, F.; Nadeau, V.; Tremblay, E.; et al. Potassium Channel Subfamily K Member 3 (KCNK3) Contributes to the Development of Pulmonary Arterial Hypertension. Circulation 2016, 133, 1371–1385. [Google Scholar] [CrossRef]
  263. Kitagawa, M.G.; Reynolds, J.O.; Wehrens, X.H.; Bryan, R.M., Jr.; Pandit, L.M. Hemodynamic and Pathologic Characterization of the TASK-1(-/-) Mouse Does Not Demonstrate Pulmonary Hypertension. Front. Med. (Lausanne) 2017, 4, 177. [Google Scholar] [CrossRef] [Green Version]
  264. Tang, G.; Wu, L.; Wang, R. The Effect of Hydroxylamine on KATP Channels in Vascular Smooth Muscle and Underlying Mechanisms. Mol. Pharmacol. 2005, 67, 1723–1731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Hattori, T.; Kajikuri, J.; Katsuya, H.; Itoh, T. Effects of H2O2 on membrane potential of smooth muscle cells in rabbit mesenteric resistance artery. Eur. J. Pharmacol. 2003, 464, 101–109. [Google Scholar] [CrossRef]
  266. Wei, E.P.; Kontos, H.A.; Beckman, J.S. Mechanisms of cerebral vasodilation by superoxide, hydrogen peroxide, and peroxynitrite. Am. J. Physiol. 1996, 271 Pt 2, H1262–H1266. [Google Scholar] [CrossRef]
  267. Fukumoto, M.; Nakaizumi, A.; Zhang, T.; Lentz, S.I.; Shibata, M.; Puro, D.G. Vulnerability of the retinal microvasculature to oxidative stress: Ion channel-dependent mechanisms. Am. J. Physiol. Cell Physiol. 2012, 302, C1413–C1420. [Google Scholar] [CrossRef] [PubMed]
  268. Ichinari, K. Direct Activation of the ATP-sensitive Potassium Channel by Oxygen Free Radicals in Guinea-pig Ventricular Cells:its Potentiation by MgADP. J. Mol. Cell. Cardiol. 1996, 28, 1867–1877. [Google Scholar] [CrossRef]
  269. Ohashi, M.; Faraci, F.; Heistad, D. Peroxynitrite hyperpolarizes smooth muscle and relaxes internal carotid artery in rabbit via ATP-sensitive K+ channels. Am. J. Physiol. Heart Circ. Physiol. 2005, 289, H2244–H2250. [Google Scholar] [CrossRef]
  270. Armstead, W.M. Endothelin-Induced Cyclooxygenase-Dependent Superoxide Generation Contributes to K+ Channel Functional Impairment after Brain Injury. J. Neurotrauma 2001, 18, 1039–1048. [Google Scholar] [CrossRef]
  271. Yasui, S.; Mawatari, K.; Morizumi, R.; Furukawa, H.; Shimohata, T.; Harada, N.; Takahashi, A.; Nakaya, Y. Hydrogen peroxide inhibits insulin-induced ATP-sensitive potassium channel activation independent of insulin signaling pathway in cultured vascular smooth muscle cells. J. Med Investig. 2012, 59, 36–44. [Google Scholar] [CrossRef] [Green Version]
  272. Thengchaisri, N.; Kuo, L. Hydrogen peroxide induces endothelium-dependent and -independent coronary arteriolar dilation: Role of cyclooxygenase and potassium channels. Am. J. Physiol. Heart Circ. Physiol. 2003, 285, H2255–H2263. [Google Scholar] [CrossRef] [Green Version]
  273. Zhang, D.X.; Borbouse, L.; Gebremedhin, D.; Mendoza, S.A.; Zinkevich, N.S.; Li, R.; Gutterman, D.D. H2O2-induced dilation in human coronary arterioles: Role of protein kinase G dimerization and large-conductance Ca2+-activated K+ channel activation. Circ. Res. 2012, 110, 471–480. [Google Scholar] [CrossRef] [Green Version]
  274. Hayabuchi, Y.; Nakaya, Y.; Matsuoka, S.; Kuroda, Y. Hydrogen peroxide-induced vascular relaxation in porcine coronary arteries is mediated by Ca2+-activated K+ channels. Heart Vessel. 1998, 13, 9–17. [Google Scholar] [CrossRef] [PubMed]
  275. Iida, Y.; Katusic, Z.S.; Wei, E.P. Mechanisms of Cerebral Arterial Relaxations to Hydrogen Peroxide. Stroke 2000, 31, 2224–2230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Zhao, K.S.; Liu, J.; Yang, G.Y.; Jin, C.; Huang, Q.; Huang, X. Peroxynitrite leads to arteriolar smooth muscle cell membrane hyperpolarization and low vasoreactivity in severe shock. Clin. Hemorheol. Microcirc. 2000, 23, 259–267. [Google Scholar] [PubMed]
  277. Au, A.L.; Seto, S.W.; Chan, S.; Chan, M.; Kwan, Y. Modulation by homocysteine of the iberiotoxin-sensitive, Ca2+-activated K+ channels of porcine coronary artery smooth muscle cells. Eur. J. Pharmacol. 2006, 546, 109–119. [Google Scholar] [CrossRef]
  278. Brakemeier, S.; Eichler, I.; Knorr, A.; Fassheber, T.; Köhler, R.; Hoyer, J.; Köhler, R. Modulation of Ca2+-activated K+ channel in renal artery endothelium in situ by nitric oxide and reactive oxygen species. Kidney Int. 2003, 64, 199–207. [Google Scholar] [CrossRef] [Green Version]
  279. Liu, Y.; Terata, K.; Chai, Q.; Li, H.; Kleinman, L.H.; Gutterman, D.D. Peroxynitrite inhibits Ca2+-activated K+ channel activity in smooth muscle of human coronary arterioles. Circ. Res. 2002, 91, 1070–1076. [Google Scholar] [CrossRef] [Green Version]
  280. Frisbee, J.C.; Maier, K.G.; Stepp, D.W. Oxidant stress-induced increase in myogenic activation of skeletal muscle resistance arteries in obese Zucker rats. Am. J. Physiol. Heart Circ. Physiol. 2002, 283, H2160–H2168. [Google Scholar] [CrossRef] [Green Version]
  281. Somlyo, A.P.; Somlyo, A.V. Ca2+ Sensitivity of Smooth Muscle and Nonmuscle Myosin II: Modulated by G Proteins, Kinases, and Myosin Phosphatase. Physiol. Rev. 2003, 83, 1325–1358. [Google Scholar] [CrossRef] [Green Version]
  282. Fagan, K.A.; Oka, M.; Bauer, N.R.; Gebb, S.A.; Ivy, D.D.; Morris, K.G.; McMurtry, I.F. Attenuation of acute hypoxic pulmonary vasoconstriction and hypoxic pulmonary hypertension in mice by inhibition of Rho-kinase. Am. J. Physiol. Lung Cell. Mol. Physiol. 2004, 287, L656–L664. [Google Scholar] [CrossRef]
  283. Kitazawa, T.; Eto, M.; Woodsome, T.P.; Brautigan, D.L. Agonists Trigger G Protein-mediated Activation of the CPI-17 Inhibitor Phosphoprotein of Myosin Light Chain Phosphatase to Enhance Vascular Smooth Muscle Contractility. J. Biol. Chem. 2000, 275, 9897–9900. [Google Scholar] [CrossRef] [Green Version]
  284. Hartmann, S.; Ridley, A.J.; Lutz, S. The Function of Rho-Associated Kinases ROCK1 and ROCK2 in the Pathogenesis of Cardiovascular Disease. Front. Pharmacol. 2015, 6, 276. [Google Scholar] [CrossRef]
  285. Pankey, E.A.; Byun, R.J.; Smith, W.B.; Bhartiya, M.; Bueno, F.R.; Badejo, A.M.; Stasch, J.-P.; Murthy, S.N.; Nossaman, B.; Kadowitz, P.J. The Rho kinase inhibitor azaindole-1 has long-acting vasodilator activity in the pulmonary vascular bed of the intact chest rat. Can. J. Physiol. Pharmacol. 2012, 90, 825–835. [Google Scholar] [CrossRef] [PubMed]
  286. Abe, K.; Tawara, S.; Oi, K.; Hizume, T.; Uwatoku, T.; Fukumoto, Y.; Kaibuchi, K.; Shimokawa, H. Long-Term Inhibition of Rho-kinase Ameliorates Hypoxia-Induced Pulmonary Hypertension in Mice. J. Cardiovasc. Pharmacol. 2006, 48, 280–285. [Google Scholar] [CrossRef] [PubMed]
  287. Oka, M.; A Fagan, K.; Jones, P.L.; McMurtry, I.F. Therapeutic potential of RhoA/Rho kinase inhibitors in pulmonary hypertension. Br. J. Pharmacol. 2008, 155, 444–454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  288. Guilluy, C.; Sauzeau, V.; Rolli-Derkinderen, M.; Guerin, P.; Sagan, C.; Pacaud, P.; Loirand, G. Inhibition of RhoA/Rho kinase pathway is involved in the beneficial effect of sildenafil on pulmonary hypertension. Br. J. Pharmacol. 2005, 146, 1010–1018. [Google Scholar] [CrossRef] [Green Version]
  289. Ziino, A.J.; Ivanovska, J.; Belcastro, R.; Kantores, C.; Xu, E.Z.; Lau, M.; McNamara, P.J.; Tanswell, A.K.; Jankov, R.P. Effects of rho-kinase inhibition on pulmonary hypertension, lung growth, and structure in neonatal rats chronically exposed to hypoxia. Pediatr. Res. 2010, 67, 177–182. [Google Scholar] [CrossRef] [Green Version]
  290. Wang, Z.; Lannér, M.C.; Jin, N.; Swartz, D.; Li, L.; Rhoades, R.A. Hypoxia inhibits myosin phosphatase in pulmonary arterial smooth muscle cells: Role of Rho-kinase. Am. J. Respir. Cell Mol. Biol. 2003, 29, 465–471. [Google Scholar] [CrossRef] [Green Version]
  291. Woodsome, T.P.; Polzin, A.; Kitazawa, K.; Eto, M. Agonist- and depolarization-induced signals for myosin light chain phosphorylation and force generation of cultured vascular smooth muscle cells. J. Cell Sci. 2006, 119 Pt 9, 1769–1780. [Google Scholar] [CrossRef] [Green Version]
  292. Mita, M.; Tanaka, H.; Yanagihara, H.; Nakagawà, J.-I.; Hishinuma, S.; Sutherland, C.; Walsh, M.P.; Shoji, M. Membrane depolarization-induced RhoA/Rho-associated kinase activation and sustained contraction of rat caudal arterial smooth muscle involves genistein-sensitive tyrosine phosphorylation. J. Smooth Muscle Res. 2013, 49, 26–45. [Google Scholar] [CrossRef] [Green Version]
  293. Porras-González, C.; Ordoñez, A.; Castellano, A.; Ureña, J. Regulation of RhoA/ROCK and sustained arterial contraction by low cytosolic Ca2+ levels during prolonged depolarization of arterial smooth muscle. Vasc. Pharmacol. 2017, 93, 33–41. [Google Scholar] [CrossRef]
  294. Sakurada, S.; Takuwa, N.; Sugimoto, N.; Wang, Y.; Seto, M.; Sasaki, Y.; Takuwa, N. Ca2+-dependent activation of Rho and Rho kinase in membrane depolarization-induced and receptor stimulation-induced vascular smooth muscle contraction. Circ. Res. 2003, 93, 548–556. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  295. Fukata, Y.; Kaibuchi, K.; Amano, M. Rho–Rho-kinase pathway in smooth muscle contraction and cytoskeletal reorganization of non-muscle cells. Trends Pharmacol. Sci. 2001, 22, 32–39. [Google Scholar] [CrossRef]
  296. Turcotte, S.; Desrosiers, R.R.; Béliveau, R. HIF-1alpha mRNA and protein upregulation involves Rho GTPase expression during hypoxia in renal cell carcinoma. J. Cell Sci. 2003, 116 Pt 11, 2247–2260. [Google Scholar] [CrossRef] [Green Version]
  297. Wennerberg, K.; Der, C.J. Rho-family GTPases: It’s not only Rac and Rho (and I like it). J. Cell Sci. 2004, 117 Pt 8, 1301–1312. [Google Scholar] [CrossRef] [Green Version]
  298. Heo, J.; Campbell, S.L. Mechanism of Redox-mediated Guanine Nucleotide Exchange on Redox-active Rho GTPases. J. Biol. Chem. 2005, 280, 31003–31010. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  299. Aghajanian, A.; Wittchen, E.S.; Campbell, S.L.; Burridge, K. Direct Activation of RhoA by Reactive Oxygen Species Requires a Redox-Sensitive Motif. PLoS ONE 2009, 4, e8045. [Google Scholar] [CrossRef] [Green Version]
  300. Chaumais, M.-C.; Ranchoux, B.; Montani, D.; Dorfmüller, P.; Tu, L.; Lecerf, F.; Raymond, N.; Guignabert, C.; Price, L.; Simonneau, G.; et al. N-acetylcysteine improves established monocrotaline-induced pulmonary hypertension in rats. Respir. Res. 2014, 15, 65. [Google Scholar] [CrossRef] [Green Version]
  301. He, B.; Zhang, Y.; Kang, B.; Xiao, J.; Xie, B.; Wang, Z. Protection of oral hydrogen water as an antioxidant on pulmonary hypertension. Mol. Biol. Rep. 2013, 40, 5513–5521. [Google Scholar] [CrossRef] [Green Version]
  302. Jefferson, J.A.; Simoni, J.M.; Escudero, E.; Hurtado, M.-E.; Swenson, E.R.; E Wesson, D.; Schreiner, G.F.; Schoene, R.B.; Johnson, R.J.; Hurtado, A. Increased Oxidative Stress Following Acute and Chronic High Altitude Exposure. High Alt. Med. Biol. 2004, 5, 61–69. [Google Scholar] [CrossRef] [PubMed]
  303. Cracowski, J.-L.; Cracowski, C.; Bessard, G.; Pépin, J.-L.; Bessard, J.; Schwebel, C.; Stanke-Labesque, F.; Pison, C. Increased Lipid Peroxidation in Patients with Pulmonary Hypertension. Am. J. Respir. Crit. Care Med. 2001, 164, 1038–1042. [Google Scholar] [CrossRef] [PubMed]
  304. Smukowska-Gorynia, A.D.; Rzymski, P.; Marcinkowska, J.; Poniedziałek, B.; Komosa, A.; Cieslewicz, A.; Slawek-Szmyt, S.; Janus, M.; Araszkiewicz, A.; Jankiewicz, S.; et al. Prognostic Value of Oxidative Stress Markers in Patients with Pulmonary Arterial or Chronic Thromboembolic Pulmonary Hypertension. Oxid. Med. Cell. Longev. 2019, 2019, 3795320. [Google Scholar] [CrossRef] [PubMed]
  305. Robbins, I.M.; Morrow, J.D.; Christman, B.W. Oxidant stress but not thromboxane decreases with epoprostenol therapy. Free Radic. Biol. Med. 2005, 38, 568–574. [Google Scholar] [CrossRef]
  306. Hemnes, A.R.; Rathinasabapathy, A.; Austin, E.A.; Brittain, E.L.; Carrier, E.J.; Chen, X.; Fessel, J.P.; Fike, C.D.; Fong, P.; Fortune, N.; et al. A potential therapeutic role for angiotensin-converting enzyme 2 in human pulmonary arterial hypertension. Eur. Respir. J. 2018, 51, 1702638. [Google Scholar] [CrossRef]
  307. Westerhof, B.E.; Saouti, N.; Van Der Laarse, W.J.; Westerhof, N.; Noordegraaf, A.V. Treatment strategies for the right heart in pulmonary hypertension. Cardiovasc. Res. 2017, 113, 1465–1473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  308. Sharp, J.; Farha, S.; Park, M.M.; Comhair, S.A.; Lundgrin, E.L.; Tang, W.W.; Bongard, R.D.; Merker, M.P.; Erzurum, S.C. Coenzyme Q supplementation in pulmonary arterial hypertension. Redox Biol. 2014, 2, 884–891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Enhanced vasoconstriction resulting from chronic hypoxia-induced functional alterations of endothelial and smooth muscle cells contributes to pulmonary hypertension. PAEC, pulmonary arterial endothelial cell; PASMC, pulmonary arterial smooth muscle cell; ROS, reactive oxygen species; MLCK, myosin light chain kinase; MLCP, myosin light chain phosphatase; MLC, myosin light chain.
Figure 1. Enhanced vasoconstriction resulting from chronic hypoxia-induced functional alterations of endothelial and smooth muscle cells contributes to pulmonary hypertension. PAEC, pulmonary arterial endothelial cell; PASMC, pulmonary arterial smooth muscle cell; ROS, reactive oxygen species; MLCK, myosin light chain kinase; MLCP, myosin light chain phosphatase; MLC, myosin light chain.
Antioxidants 09 00999 g001
Figure 2. O2.− is generated from various enzymatic sources and converted to other forms of ROS, including H2O2 and ONOO. H2O2 can also be produced from NOX4 and MAOs. PAEC, pulmonary arterial endothelial cell; PASMC, pulmonary arterial smooth muscle cell; ROS, reactive oxygen species; O2.−, superoxide; H2O2, hydrogen peroxide; ONOO, peroxynitrite; H2O, water; NOX, NADPH oxidase; eNOS, endothelial nitric oxide synthase; XO, xanthine oxidase; MAO, monoamine oxidase; SOD, superoxide dismutase.
Figure 2. O2.− is generated from various enzymatic sources and converted to other forms of ROS, including H2O2 and ONOO. H2O2 can also be produced from NOX4 and MAOs. PAEC, pulmonary arterial endothelial cell; PASMC, pulmonary arterial smooth muscle cell; ROS, reactive oxygen species; O2.−, superoxide; H2O2, hydrogen peroxide; ONOO, peroxynitrite; H2O, water; NOX, NADPH oxidase; eNOS, endothelial nitric oxide synthase; XO, xanthine oxidase; MAO, monoamine oxidase; SOD, superoxide dismutase.
Antioxidants 09 00999 g002
Figure 3. ROS facilitate pulmonary arterial constriction in pulmonary hypertension by make various posttranslational modifications at cysteine residues of ion channels and molecules.
Figure 3. ROS facilitate pulmonary arterial constriction in pulmonary hypertension by make various posttranslational modifications at cysteine residues of ion channels and molecules.
Antioxidants 09 00999 g003
Figure 4. Summary of Ca2+-dependent influx and release mechanisms in pulmonary arterial smooth muscle cells following chronic hypoxia. See text for details. KV, voltage-gated K+ channel; VGCC, voltage-gated Ca2+ channel; SOC, store-operated channel; ROC, receptor-operated channel; MSC, mechanosensitive channel; GPCR, G protein-coupled receptor; PLC, phospholipase C; PIP2, phosphatidylinositol 4,5-bisphosphate; DAG, diacylglycerol; IP3, inositol triphosphate; SR, sarcoplasmic reticulum; MLCK, myosin light chain kinase; MLCP, myosin light chain phosphatase; MLC, myosin light chain.
Figure 4. Summary of Ca2+-dependent influx and release mechanisms in pulmonary arterial smooth muscle cells following chronic hypoxia. See text for details. KV, voltage-gated K+ channel; VGCC, voltage-gated Ca2+ channel; SOC, store-operated channel; ROC, receptor-operated channel; MSC, mechanosensitive channel; GPCR, G protein-coupled receptor; PLC, phospholipase C; PIP2, phosphatidylinositol 4,5-bisphosphate; DAG, diacylglycerol; IP3, inositol triphosphate; SR, sarcoplasmic reticulum; MLCK, myosin light chain kinase; MLCP, myosin light chain phosphatase; MLC, myosin light chain.
Antioxidants 09 00999 g004
Figure 5. ROS modulation of Ca2+ influx. See text for details. L-type VGCC, L-type voltage-gated Ca2+ channel; TRPC, canonical transient receptor potential channel; TRPV4, transient receptor potential vanilloid 4; ASIC 1, acid sensing ion channel 1; MSC, mechanosensitive channel; NOX2, NADPH oxidase 2; PLC, phospholipase C; PIP2, phosphatidylinositol 4,5-bisphosphate; DAG, diacylglycerol; PKC, protein kinase C; SR, sarcoplasmic reticulum; STIM1, stromal interaction molecule 1; ROS, reactive oxygen species; O2.−, superoxide; H2O2, hydrogen peroxide; ONOO, peroxynitrite; SOD, superoxide dismutase; NO, nitric oxide; Fyn, Fyn kinase; GSSG, oxidized glutathione; GSH, reduced glutathione; DTT, dithiothreitol; DTNB, 5,5′-Dithiobis(2-nitrobenzoic acid).
Figure 5. ROS modulation of Ca2+ influx. See text for details. L-type VGCC, L-type voltage-gated Ca2+ channel; TRPC, canonical transient receptor potential channel; TRPV4, transient receptor potential vanilloid 4; ASIC 1, acid sensing ion channel 1; MSC, mechanosensitive channel; NOX2, NADPH oxidase 2; PLC, phospholipase C; PIP2, phosphatidylinositol 4,5-bisphosphate; DAG, diacylglycerol; PKC, protein kinase C; SR, sarcoplasmic reticulum; STIM1, stromal interaction molecule 1; ROS, reactive oxygen species; O2.−, superoxide; H2O2, hydrogen peroxide; ONOO, peroxynitrite; SOD, superoxide dismutase; NO, nitric oxide; Fyn, Fyn kinase; GSSG, oxidized glutathione; GSH, reduced glutathione; DTT, dithiothreitol; DTNB, 5,5′-Dithiobis(2-nitrobenzoic acid).
Antioxidants 09 00999 g005
Figure 6. Summary of Ca2+ sensitization in pulmonary arterial smooth muscle cells. Myofilament Ca2+ sensitization is facilitated by ROS following CH. In particular, membrane stretch and endothelin 1 (ET-1) activate Src kinase-epidermal growth factor receptor (EGFR)-NADPH oxidase 2 (NOX2) signaling axis that contributes to CH-induced augmentation of Ca2+-independent pulmonary vasoconstriction and pulmonary hypertension. See text for details. GPCR, G protein-coupled receptor; PLC, phospholipase C; PIP2, phosphatidylinositol 4,5-bisphosphate; DAG, diacylglycerol; PKC, protein kinase C; O2.−, superoxide; ROK, Rho kinase; MLCP, myosin light chain phosphatase; MLCK, myosin light chain kinase; MLC, myosin light chain.
Figure 6. Summary of Ca2+ sensitization in pulmonary arterial smooth muscle cells. Myofilament Ca2+ sensitization is facilitated by ROS following CH. In particular, membrane stretch and endothelin 1 (ET-1) activate Src kinase-epidermal growth factor receptor (EGFR)-NADPH oxidase 2 (NOX2) signaling axis that contributes to CH-induced augmentation of Ca2+-independent pulmonary vasoconstriction and pulmonary hypertension. See text for details. GPCR, G protein-coupled receptor; PLC, phospholipase C; PIP2, phosphatidylinositol 4,5-bisphosphate; DAG, diacylglycerol; PKC, protein kinase C; O2.−, superoxide; ROK, Rho kinase; MLCP, myosin light chain phosphatase; MLCK, myosin light chain kinase; MLC, myosin light chain.
Antioxidants 09 00999 g006
Table 1. Expression of NOX isoforms in the pulmonary vasculature.
Table 1. Expression of NOX isoforms in the pulmonary vasculature.
NOX IsoformExpression in PAECExpression in PASMCROS Generated
NOX1Human [81,82,83], rat [84]Human [85], rat [86,87], mouse [85,88]O2.− [83,85,86,87,88]
NOX2Human [81,82,89,90], rat [84], mouse [89,91]Rat [92]O2.− [81,89]
NOX3Human [82]N/AO2.− [93]
NOX4Human [81,82,94], rat [84], mouse [95]Human [90,96], rat [86,92], mouse [90]O2.− [82,97], H2O2 [82,95,96]
Table 2. Ca2+ influx in CH-induced PH.
Table 2. Ca2+ influx in CH-induced PH.
Channel/MoleculeAlteration by CHFunctions
L-type VGCCIncreased current density [195], channel upregulation (Cav1.2) [196]Positive:
Mediate CH-induced enhanced pulmonary vascular tone [195] and PA vasoconstriction to KCl [195,196] and to L-type VGCC activator [195]
Negative:
1. Without effects on basal [Ca2+]i in PASMCs or basal PA tension [180]
2. Responsible for 30–40% of basal [Ca2+]i in cultured PASMCs but not affect basal PA tone [178]
3. Do not contribute to increase PA wall basal [Ca2+]i or elevated PA constriction to UTP following CH [17]
4. Do not contribute to CH-induced augmentation of PA myogenic tone [21]
5. CH-induced PH is not acutely alleviated by L-type VGCC inhibition in SD rats [197] or COPD patients [198,199]
T-type VGCCChannel upregulation (Cav3.2) [196]Positive:
Mediate CH-induced augmented PA constriction to K+ and U-46619 [196]
Negative:
1. Do not contribute to increase PA wall basal [Ca2+]i following CH [17]
2. Do not contribute to CH-induced augmentation of PA myogenic tone [21]
TRPC1Channel upregulation [177,178,200]CH-induced PH [201,202]; SOCE in PASMC [178,200]; CH-induced augmented basal tone and vasoconstriction to 5-HT [202]
TRPC6Channel upregulation [177,178,203]CH-induced PH [202,203]; ROCE in PASMC [178]; augmented SOCE in PASMCs following CH [203]; basal tone under normoxia [202]; CH-induced augmented vasoconstriction to 5-HT [202]
TRPV4Channel upregulation in PASMCs [204,205], increased channel activities in PASMCs [204,205]1. CH-induced PH development [204,206]
2. CH-induced enhanced myogenic tone [204] and augmented vasoconstriction to serotonin [206] and TRPV4 agonist [205] but not to U46619 [204], PE [206] or ET-1 [206] in endothelium-disrupted PAs
3. Ca2+-induced Ca2+ release in PASMCs [205]
ASIC1Unaltered expression [207]Contribute to augmented SOCE and SOCE-induced vasoconstriction in PAs following CH [17]; CH-induced PH [207]
Orai1Upregulation [14,200,203,208]CH-induced increases in basal Ca2+ [14] and SOCE [14,200] in PASMCs
Orai2Upregulation [14,203,208]CH-induced increases in basal Ca2+ and SOCE in PASMCs [14]
Orai3Unaltered expression [14]CH-induced increases in basal Ca2+ and SOCE in PASMCs [14]
STIM1Upregulation [200,208], unaltered expression [14]CH-induced increases in basal Ca2+ [14] and SOCE [14,200] in PASMCs
STIM2Upregulation [203,208]Enhanced SOCE in PASMCs from PH patients [209]
MSCIncreased channel activities [176]CH-induced augmentation of PA myogenic tone [19,21,176]
Table 4. Other K+ channels involved in CH-induced PH or PAH.
Table 4. Other K+ channels involved in CH-induced PH or PAH.
TypeChannelFunctionRef.
KirKATPGain of function protects against CH-induced PH indices including mPAP, RV hypertrophy and PA remodeling[256]
KCaLarge conductance KCa (BKCa)Gain of function protects against monocrotaline-induced PH, reduces PDGF-induced PASMC proliferation[257]
Loss of function does not affect PH development following CH[258]
K2PTREK-1 (K2P2.1)Gain of function leads to PAEC hyperpolarization and PA relaxation[259]
K2PTWIK-2 (K2P6.1)Gain of function leads to PAEC hyperpolarization and PA relaxation[259]
K2PTWIK-2 (KCNK6)Loss of function results in increased RVSP, PA thickening, greater PA vasoconstrictor to U46619[260]
Loss of function causes PASMC depolarization, enhanced [Ca2+]i and PA constriction to U46619[261]
K2PTASK-1 (KCNK3)Loss of function favors proliferation of (PAEC, PASMC and fibroblast) and enhanced basal tone[262]
Gain of function protects against monocrotaline-induced PH[262]
Loss of function is without effects on CH-induced PH[263]
Table 5. ROS modulation of KATP and KCa channels in cardiovascular system.
Table 5. ROS modulation of KATP and KCa channels in cardiovascular system.
OutcomeROS
O2.−H2O2ONOO
KATP activationMesenteric artery SMC [264]Mesenteric arteries [265]
Cerebral arteries [266]
Retinal microvessels [267]
Cardiomyocytes [268]
Cerebral arteries [266]
Internal carotid arteries [269]
KATP inhibitionCerebral arteries [270]A10 cell line [271]N/A
KCa activationCerebral arteries [266]Coronary arteries [272,273,274]
Cerebral arteries [266,275]
Arteriolar SMC [276]
KCa inhibitionCoronary arteries [277] Cerebral arteries [270]Renal arteries [278]Coronary artery SMC [279]
Gracilis arteries [280]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yan, S.; Resta, T.C.; Jernigan, N.L. Vasoconstrictor Mechanisms in Chronic Hypoxia-Induced Pulmonary Hypertension: Role of Oxidant Signaling. Antioxidants 2020, 9, 999. https://doi.org/10.3390/antiox9100999

AMA Style

Yan S, Resta TC, Jernigan NL. Vasoconstrictor Mechanisms in Chronic Hypoxia-Induced Pulmonary Hypertension: Role of Oxidant Signaling. Antioxidants. 2020; 9(10):999. https://doi.org/10.3390/antiox9100999

Chicago/Turabian Style

Yan, Simin, Thomas C. Resta, and Nikki L. Jernigan. 2020. "Vasoconstrictor Mechanisms in Chronic Hypoxia-Induced Pulmonary Hypertension: Role of Oxidant Signaling" Antioxidants 9, no. 10: 999. https://doi.org/10.3390/antiox9100999

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop