Next Article in Journal
Bioactive Activities of the Phenolic Extract from Sterile Bracts of Araucaria angustifolia
Previous Article in Journal
Correlation between Blunted Nocturnal Decrease in Diastolic Blood Pressure and Oxidative Stress: An Observational Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Oxidative Regulation of Vascular Cav1.2 Channels Triggers Vascular Dysfunction in Hypertension-Related Disorders

Lawrence D. Longo MD Center for Perinatal Biology, Department of Basic Sciences, Loma Linda University School of Medicine, Loma Linda, CA 92350, USA
*
Authors to whom correspondence should be addressed.
Antioxidants 2022, 11(12), 2432; https://doi.org/10.3390/antiox11122432
Submission received: 8 November 2022 / Revised: 28 November 2022 / Accepted: 6 December 2022 / Published: 9 December 2022

Abstract

:
Blood pressure is determined by cardiac output and peripheral vascular resistance. The L-type voltage-gated Ca2+ (Cav1.2) channel in small arteries and arterioles plays an essential role in regulating Ca2+ influx, vascular resistance, and blood pressure. Hypertension and preeclampsia are characterized by high blood pressure. In addition, diabetes has a high prevalence of hypertension. The etiology of these disorders remains elusive, involving the complex interplay of environmental and genetic factors. Common to these disorders are oxidative stress and vascular dysfunction. Reactive oxygen species (ROS) derived from NADPH oxidases (NOXs) and mitochondria are primary sources of vascular oxidative stress, whereas dysfunction of the Cav1.2 channel confers increased vascular resistance in hypertension. This review will discuss the importance of ROS derived from NOXs and mitochondria in regulating vascular Cav1.2 and potential roles of ROS-mediated Cav1.2 dysfunction in aberrant vascular function in hypertension, diabetes, and preeclampsia.

Graphical Abstract

1. Introduction

Free intracellular Ca2+ functions as an important second messenger to regulate various physiological processes including muscle contractility, gene expression, hormone secretion, and neuronal transmission [1,2]. The Ca2+ signal transduction is initiated by increasing intracellular Ca2+ concentration ([Ca2+]i) due to Ca2+ entry through ion channels on the cell membrane and Ca2+ release via Ca2+-release channels on the endo/sarcoplasmic reticulum. Ca2+ influx in vascular smooth muscle cells is primarily mediated by L-type voltage-gated Ca2+ (Cav1.2) channels encoded by the CACNA1C gene [3]. Vascular tone in small arteries and arterioles, a major determinant of blood pressure and blood flow, is governed by the contractile state of vascular smooth muscle cells which is determined by dynamic changes in intracellular [Ca2+]i [4]. An increase in [Ca2+]i instigates vascular smooth muscle cell contraction, whereas a decrease in [Ca2+]i promotes relaxation.
Elevated blood pressure is a prime characteristic of hypertension and preeclampsia [5,6,7]. Preeclampsia is a multisystem disorder characterized by the onset of hypertension after 20 weeks of gestation. In addition, there is a strong epidemiologic association of diabetes and hypertension, and hypertension is common among diabetic patients [8,9,10]. Prevalence of hypertension in adults is ~30% worldwide in 2010 and in United States between 2011–2012 [11,12,13]. Remarkably, prevalence of hypertension in adults with diabetes is 66–76% in United States [10]. Preeclampsia affects ~5% of pregnancies globally [14]. Hereafter, these three disorders are collectively referred as hypertension-related disorders in this review. Hypertension is categorized into essential (primary) and secondary hypertension. Essential hypertension accounts for 95% human cases and has no known cause, whereas the remaining are secondary hypertension which occurs due to known diseases or conditions including renovascular disease, aldosteronism, Cushing’s syndrome, thyroid dysfunction, and among others [15,16]. Diabetes is commonly classified into two types. Type 1 diabetes results from insulin deficiency due to autoimmune destruction of β-cells in the pancreas, whereas type 2 diabetes originates from insulin resistance owing to reduced insulin sensitivity and accounts for >90% of all cases [17,18]. Hypertension is the leading cause of cardiovascular disease worldwide, including stroke, coronary artery disease, heart failure, arrhythmia, and peripheral vascular disease [19].
High blood pressure arises from vascular dysfunction [20]. Oxidative stress is a hallmark of hypertension, preeclampsia, and diabetes [21,22,23]. Excessive reactive oxygen species (ROS) play an important role in vascular dysfunction and contribute to the pathogenesis of hypertension-related disorders [24,25,26]. Given the pivotal role of Cav1.2 in regulating vascular tone, we discuss in this review the oxidative modification leading to alternations of vascular Cav1.2 function/expression and enhanced myogenic tone in hypertension-related disorders.

2. Cav1.2 in Vascular Smooth Muscle

2.1. Overview of Cav1.2

L-type Ca2+ (Cav) channels are heteromultimeric complexes comprising pore-forming α1c and auxiliary β, α2δ, and γ subunits [27]. The α1c subunit possesses four repeat domains (I–IV) linked by intracellular loops and intracellular NH2-/COOH-termini with each domain containing six transmembrane segments (S1–S6) (Figure 1). The ion-conducting pore is formed by S5 and S6 and the loop between them, whereas the voltage sensor is located in the S4 segments [28]. The intracellular COOH terminus, along with intracellular loops, plays important roles in Ca2+-dependent inactivation, channel trafficking, phosphorylation and oxidation [28,29,30]. Cav is activated by membrane depolarization and its activation allows Ca2+ influx through the channel pore. The expression of the α1c subunit itself can form functional channels to conduct Ca2+ ions, whereas the incorporation of auxiliary subunits promotes membrane α1c expression and alters biophysical properties of the channel [28,31]. Cav channels are sensitive to blockades by dihydropyridines (i.e., nifedipine), phenylalkylamines (i.e., verapamil) and benzothiazepines (i.e., diltiazem) [32]. There are four types of Cav channels (Cav1.1–1.4). Cav1.2 is predominantly expressed in the heart and in vascular smooth muscle [33,34]. Ca2+ influx through the channel is a primary trigger of vasoconstriction and also participates in transcriptional regulation. The auxiliary subunits β2, β3, and α2δ1 are expressed in vascular smooth muscle [35,36,37,38]. β subunits, lacking membrane-spanning segments, are located intracellularly and interact with the α interaction domain (AID) on the I-II linker of the α1c subunit [39]. The α2δ1 subunit is a single gene product bound together by disulfide bonds. Whereas the α2 subunit is extracellular, the δ subunit contains a single transmembrane segment [40]. The diversity of Cav1.2 is also conferred by alternative splicing. Smooth muscle contains Cav1.2 exons 1, 8, 9 *, 31/32, and 33 [41,42]. Due to the scope of this review, only a brief description of Cav1.2 is presented here. The reader is referred to the literature for detailed information on Cav1.2 [43,44,45].

2.2. Regulation of Cav1.2

2.2.1. Regulation by Auxiliary Subunits

The molecular composition of Cav1.2 in vascular smooth muscle cells includes the pore-forming α1c subunit and auxiliary β2/3 and α2/δ1 subunits. The β3 subunit appears to be the principal β isoform in vascular smooth muscle cells [35]. Genetic deletion of the β3 subunit resulted in reduced α1c expression in mouse aorta and is associated with a reduction in Ca2+ channel current and a slower inactivation rate [35]. This genetic manipulation also attenuates angiotensin II-induced upregulation of Cav1.2 channels in mouse mesenteric arteries and the development of hypertension [38]. Similarly, knocking-down α2/δ1 with short hairpin RNA in cerebral artery reduced plasma membrane α1c expression [36]. Moreover, Cav1.2 currents were inhibited by pregabalin, an α2/δ1/2 ligand, and by an α2/δ-1 antibody in cerebral vascular smooth muscle cells, leading to vasodilation [36].

2.2.2. Regulation by Protein Kinases

Protein phosphorylation, a covalent addition of the phosphate group to the side chain of serine, threonine, and tyrosine residues by protein kinases, is a common posttranslational modification to fine tune activities of receptors, ion channels and enzymes. Unsurprisingly, Cav1.2 is a target of protein kinases, and its activity is subject to the regulation by protein phosphorylation. Both α1c and β2 subunits are phosphorylated by protein kinases A (PKA), C (PKC), and G (PKG) [29,46,47,48,49]. A variety of putative serine/threonine phosphorylation sites have been identified, yet their role in regulating Cav1.2 remains unsettled [50]. In vascular smooth muscle cells, the regulation of Cav1.2 by PKA is controversial. PKA is found to either inhibit or enhance Cav1.2 activity [51,52,53,54,55,56,57]. The stimulatory effect of PKA on Cav1.2 in vascular smooth muscle cells depends on anchoring adenyl cyclase 5 and PKA by A-kinase anchoring protein 150 (AKAP150) to the proximity of Cav1.2 [56,58]. Phosphorylation of Ser 1928 in the COOH-terminus of the Cav1.2 α1c subunit is required for PKA-stimulated channel activity in vascular smooth muscle cells [56]. Activation of PKG exhibits inhibitory effects on vascular Cav1.2 [52,57,59,60]. PKG mediates nitric oxide-induced inhibition of Cav1.2 [61,62]. Activation of PKC by phorbol esters and by Gq-coupled receptors also potentiates vascular Cav1.2 activity [63,64,65,66,67,68,69,70]. Basal Cav1.2 activity is evidently under the tonic control of PKC as PKC inhibition/PKCα depletion enhances Cav1.2 activity in vascular smooth muscle cells [67]. Activation of PKG by nitric oxide (NO) suppresses Cav1.2 activity [49]. Similar to PKA, PKC is anchored by AKAP150 to adjacent Cav1.2 to alter channel activity [71]. Protein tyrosine kinase c-Src promotes tyrosine phosphorylation of the α1c subunit and enhances Cav1.2 activity, which is believed to participate in regulating smooth muscle contractility [72]. c-Src via its SH2 and SH3 domains binds to the COOH-terminus of the α1c subunit [70]. Y2122 in the COOH-terminus of the α1c subunit appears to be the major phosphorylation site of c-Src [70,73]. In vascular smooth muscle cells, c-Src enhances Cav1.2 activity [64,74,75,76]. Phosphoinositide 3-kinases (PI3Ks) are found to increase Cav1.2 activity in vascular smooth muscle cells [77,78]. PI3Kγ potentiates Cav1.2 activity by facilitating plasma membrane translocation of α1c subunits and this effect is mediated by AKT/PKB-induced β2 subunit phosphorylation [79,80,81,82]. Integrins also participate in the mechanotransduction process of pressure-induced myogenic tone [83]. Integrin receptor activation is found to increase Cav1.2 activity via c-Src and PKA [84,85,86].

2.2.3. Regulation by Small GTPases

The Ras superfamily of small GTPases (also known as small G-proteins) are cellular switches that regulate a variety of biological processes in living cells. They have been implicated in regulating Ca2+ homeostasis and Cav1.2 is recognized as an important effector of the RGK subfamily (Rem, Rem2, Rad, and Gem/Kir) [87,88]. Unlike vascular Cav1.2, direct phosphorylation of the cardiac Cav1.2 α1c subunit does not contribute to PKA-stimulated channel activity as mutating all PKA consensus sites in the α1c subunit fails to block the increase in Cav1.2 activity in response to β-adrenergic stimulation [89,90]. Cardiac Cav1.2 is under tonic inhibition of Rad due to its association to both α1c and/or β subunits [91,92]. β-adrenergic stimulation induces RAD phosphorylation, promotes the release of RAD from Cav1.2, and increase channel activity [90]. Similarly, Cav1.2 activity is suppressed by other members of the RGK subfamily [93,94,95]. It remains to be determined whether vascular Cav1.2 is modulated by RGK members. Small GTPases also regulate surface expression of Cav1.2. Rab25, a member of the Rab subfamily, regulates intracellular vesicle trafficking and promotes surface expression of Cav1.2 in vascular smooth muscle cells [96].

2.3. Cav1.2 and Myogenic Tone

Vascular smooth muscle cells in resistant arteries and arterioles possess intrinsic ability to contract in response to an increase in intraluminal pressure and to relax upon a decrease in intraluminal pressure [97]. This phenomenon is defined as myogenic response and the steady-state of vascular smooth muscle cell contractility in these vessels is termed as myogenic tone. Myogenic tone sets the basal vascular tone and distribution of blood flow to and within tissues/organs. In principle, peripheral vascular resistance can be described by the Poiseuille equation: R = 8Lη/πr4, where R is the resistance, L is length of the vessel, η is viscosity of blood, and r is the radius of the vessel. According to the Poiseuille’s law, peripheral vascular resistance is inversely proportional to the radius to the fourth power. For example, decreases in vessel caliber by 10% and 20% will theoretically lead to increases in vascular resistance by ~50% and ~145%, respectively. Thus, the radius of a given small artery/arteriole has a great impact on peripheral vascular resistance and blood flow.
Vascular smooth muscle cells in resistance arteries/arterioles become depolarized in response to increased intraluminal pressure [98,99,100]. Various mechanisms have been proposed to regulate myogenic tone (Figure 2). It is widely believed that pressure-induced membrane depolarization in vascular smooth muscle cells is instigated by mechanosensitive or stretch-activated cation channels including transient receptor potential (TRP) channels and epithelial Na+ channels (ENaCs) [101,102,103,104,105,106,107]. The membrane depolarization leads to increased [Ca2+]i and vasoconstriction [99,108]. The increases in both [Ca2+]i and/or myogenic tone are blocked by the removal of extracellular Ca2+ or by Cav1.2 blockers [99,109,110,111,112,113,114,115]. These findings suggest that altered intraluminal pressure initiates myogenic tone via membrane depolarization and subsequent opening of Cav1.2. This notion is corroborated by findings from the genetic deletion of Cav1.2 α1c subunit in smooth muscle [116]. Such a manipulation abolishes myogenic reactivity in murine tibialis arteries [116]. However, Ca2+ influx mediated by T-type voltage-gated Ca2+ channels (Cav3.x) may also contribute to myogenic tone development [106,117]. Integrins, linking the extracellular matrix and cellular cytoskeleton, could also function as mechanotransducer to regulate myogenic tone via c-Src-mediated phosphorylation of Cav1.2 [83,85,118]. Moreover, some G protein-coupled receptors are mechanosensitive and produce inositol triphosphate (IP3) upon stretch to induce Ca2+ release via activating IP3R on the sarcoplasmic reticulum (SR) membrane [119,120]. Over myogenic tone, extrinsic factors such as vasodilators and vasoconstrictors can induce vasodilation and further vasoconstriction, respectively. It should be noted that the large-conductance Ca2+-activated K+ channel is also activated by the increase in intralumenal pressure, which functions as a negative feedback mechanism to limit the magnitude of intralumenal pressure-induced vasoconstriction under physiological conditions [121]. For more detailed information on the regulation of myogenic tone by ion channels, readers are referred to recent reviews on this topic [106,122,123]. As aforementioned, myogenic response is the intrinsic property of vascular smooth muscle cells to respond to changes in intraluminal vascular pressure. In general, myogenic tone development is independent of endothelium [124,125,126]. However, myogenic tone can be modified by the endothelium. For example, activation of the endothelial intermediate-conductance Ca2+-activated K+ channel is found to reduce myogenic tone [127,128]. Moreover, myogenic tone in mouse mesenteric and uterine arteries from pregnant mice is enhanced by the loss of endothelium-derived NO [129].

3. Roles of Cav1.2 in the Pathogenesis of Hypertension-Related Disorders

3.1. Aberrant Vascular Tone in Hypertension-Related Disorders

Hypertension is associated with increased peripheral vascular resistance. As myogenic tone is the fundamental element of vascular tone, it is reasonable to speculate that myogenic tone is altered in hypertension-related disorders. In patients with hypertension, myogenic tone is increased in coronary arterioles [110]. Increased myogenic tone is also noted in resistance arteries from the adipose tissue of paravertebral muscles of hypertensive patients [130]. Commonly used rat models of experimental hypertension include the spontaneously hypertensive rat (SHR), stroke-prone spontaneously hypertensive rat (SHRSP), Milan hypertensive strain (MHS), vasopressin-deficient (Di/H) rats, two-kidney, one-clip (2K1C rats), and among others. Compared to the Wistar-Kyoto rat (WKY), myogenic tone of afferent arterioles and arcuate arteries from SHR kidneys is increased [131,132]. Mesenteric arteries of SHR and MSH also exhibits higher myogenic tone than those of WKY and Milan normotensive strain (MNS), respectively [133,134]. Elevated myogenic tone is observed in cerebral arteries/basilar arteries of SHR/SHRSP/Di/H compared to WKY and Di normotensive (Di/N) rats, respectively [135,136,137,138]. Similarly, there is an increase in the myogenic tone of cremaster arterioles of SHR [139,140]. Furthermore, chronic infusion of angiotensin II in C57BL/6 mice also causes increased myogenic tone in middle cerebral arteries [141]. In contrast, mice with smooth muscle cell-specific deficiency of the mineralocorticoid receptor reduces myogenic tone and lowers blood pressure [142,143].
Myogenic tone is increased in skeletal muscle resistance arteries of diabetic patients [144]. Various animal models such as type I diabetic rodents induced by streptozotocin (STZ), obese/type II diabetic rodents induced by high-fat diet (HFD), type II diabetic HFD/STZ rodents, type II diabetic Goto-Kakizaki (GK) rats, and type II diabetic C57BL/KsJ-db/db mouse have been developed in diabetes research [145,146]. Systemic vascular resistance (also known as total vascular resistance) and arterial blood pressure are increased in C57BL/KsJ-db/db mice, HFD rats, and GK rats [147,148,149,150,151,152,153,154,155]. Mesenteric arteries of HFD mice, HFD/STZ mice, and diabetic (db/db) mice displayed higher myogenic tone compared to their counterparts [58,144,155,156]. Similarly, the cerebral arteries of STZ-Sprague Dawley rats, Bio-Breeding Zucker diabetic (BBZDR/Wor) rats, HFD mice, and HFD/STZ mice also exhibit increased myogenic tone [144,157,158]. Moreover, higher myogenic tone is detected in gracilis arterioles of STZ Wistar rats and db/db mice [148,159]. Furthermore, myogenic tone of cerebral arteries and ophthalmic arteries from C57BL/6J WT mice and Sprague Dawley rats increases in response to acute changes of glucose levels from low (10 mM) to high concentrations (20–25 mM, to mimic hyperglycemia) [58,160,161].
A study by Kublickiene and colleagues find that preeclampsia does not alter myogenic tone of myometrial arteries [162]. However, the measurement is made in only one pressure point. Nevertheless, the authors reveal that flow-mediated dilatation is lost in pressurized myometrial arteries from women with preeclampsia [162]. A frequently used rodent model of preeclampsia is induced by chronic reductions in uteroplacental perfusion (RUPP) in rats/mice by clipping/ligating the aorta below the renal arteries or the arterial and venous branches of the uterine vascular arcade [163,164]. Myogenic tone is increased in both uterine arteries and mesenteric arteries in this animal model [165,166,167]. Intriguingly, exposure to small extracellular vesicles purified from the plasma of pregnant women with preeclampsia elevates myogenic tone in the mesenteric arteries of C57Bl/6J mice [168], suggesting that circulating bioactive factor(s) could contribute to the increased peripheral vascular tone in preeclampsia. Hypoxia plays a key role in the pathogenesis of preeclampsia [169,170,171]. High-altitude hypoxia increases maternal systolic and diastolic blood pressure at term in human pregnancy [172] and is associated with increased incidence of preeclampsia [173,174,175]. Pregnant sheep exposed to normobaric hypoxia at low altitude and hypobaric hypoxia at high altitude developed preeclampsia-like symptoms including elevated maternal systemic blood pressure and increased uterine vascular resistance as well as various biochemical changes in the circulation and uteroplacental tissues [176,177,178]. Myogenic tone of resistance uterine arteries is increased in pregnant sheep acclimatized to high-altitude hypoxia [179]. This effect is replicated by ex vivo hypoxia treatment of resistance uterine arteries of low-altitude pregnant sheep [180].

3.2. Dysfunction of Vascular Cav1.2 in Hypertension-Related Disorders

As aforementioned, activation of Cav1.2 in vascular smooth muscle cells is essential for the development of myogenic tone [99,100,109,110,113,114,116]. The enhanced myogenic tone in peripheral resistance arteries and arterioles in hypertension-related disorders suggests potential dysfunction of vascular Cav1.2. Indeed, this notion is substantiated by lines of evidence from functional studies. First, the increased myogenic tone in resistance arteries/arterioles of hypertensive and diabetic animals was normalized by the Cav1.2 blocker nifedipine [58,132,181]. Whereas specific deletion of the mineralocorticoid receptor reduces KCl- and Cav1.2 agonist Bay K 8644-induced vasoconstriction of mesenteric arteries [142,182], Bay K 8644-induced contraction and increase in [Ca2+]i in 2K1C rat aorta is greater than in the control 2K rats [183]. Second, Bay K 8644 triggers greater contraction of inferior epigastric arteries of women with preeclampsia than that of normotensive subjects [184]. Similarly, Bay K 8644- or the membrane-depolarizing agent KCl-induced contraction of cerebral arteries, mesenteric arteries, renal arteries, femoral arteries is greater in SHR, diabetic STZ or GK rats, RUPP rats or pregnant rats treated with nitric oxide synthase inhibitor L-Nω-nitro arginine methyl ester (L-NAME) than their control counterparts [153,185,186,187,188,189,190,191]. Third, Bay K 8644 or KCl produces a larger increase in [Ca2+]i and contraction in renal arteries in RUPP rats and pregnant rats treated with L-NAME [186,188].
Animal models of hypertension-related disorders have provided mechanistic insights into the understanding of the Cav1.2 dysfunction in these disorders. Both aberrant expression of and dysregulation of Cav1.2 contribute to vascular Cav1.2 dysfunction. Various studies reveal increased protein expression of α1c [37,38,187,192,193], α2δ1 [37,192], and β3 [37,38] subunits in mesenteric, femoral, and cerebral arteries of SHR and angiotensin II-infused C57BL/6 mice. Consistently, increased expression of Cav1.2 is associated with enhanced channel activity in vascular smooth muscle cells [67,185,193,194,195,196,197]. The expression and activity of Cav1.2 are reduced in mesenteric arteries of aged mice lacking mineralocorticoid receptors in smooth muscle cells [143]. Dexamethasone administration increases the expression of the cardiac Cav1.2 α1c subunit in rats [198]. As expected, Cav1.2 activity in A7r5 cells is increased following chronic dexamethasone exposure [199]. Similarly, hyperthyroidism also boosts the expression of cardiac Cav1.2 [200]. Galectins (Gals) are a family of carbohydrate-binding lectins. Gal-1, through binding preferentially to Cav1.2 I–II loop without exon 9*, reduces surface expression of Cav1.2 in vascular smooth muscle cells [201]. Gal-1 is later found to compete with the β subunit for binding to the α1c subunit leading to polyubiquitination and degradation of the α1c subunit and consequent reduction of Cav1.2 activity [202]. Notably, protein levels of Gal-1 and Cav1.2 are downregulated and upregulated, respectively, in arteries of SHR and hypertensive patients [202]. In hypertensive human pulmonary arteries, the downregulation of Gal-1 is mediated by hypoxia-inducible factor 1α [202]. The knockdown of Gal-1 with siRNA increases Cav1.2 activity and promotes vasoconstriction [201], whereas the deletion of the Gal-1 encoding gene LGALS1 increases blood pressure [202].
An increase in protein expression of the α1c subunits is also observed in cerebral arteries of STZ rats, which is associated with increased Cav1.2 activity and contraction to KCl [203]. Cav1.2 also displays increased activity in vascular smooth muscle cells of cerebral and mesenteric arteries from STZ, GK, and HFD rats and db/db mice [153,204,205,206]. Human arteries from diabetic patients have higher Cav1.2 activity due to increased phosphorylation of α1C at Ser1928 by PKA [56]. Similar findings were observed in cerebral arteries of HFD mice [56]. These changes are simulated by in vitro acute exposure of rodent and human arteries to a high concentration of glucose [56,205]. A series of studies by Navedo’s group reveal that the Gs-coupled purinergic receptor P2Y11, adenylyl cyclase 5 (AC5), PKA, AKAP150, and Cav1.2 form a signaling microdomain in vascular smooth muscle cells and these components are localized in nanometer proximity within the microdomain [56,58,161]. Apparently, elevated glucose stimulates Cav1.2 activity by promoting autocrine release of nucleotides, which sequentially activates P2Y11, AC5, and PKA leading to increased phosphorylation of the α1C subunit at Ser1928 [56,58,161].
Gal-1 is highly expressed in the placenta [207]. Circulating Gal-1 level also increases during gestation [208]. Given the critical role of Gal-1 in regulating Cav1.2 surface expression/activity discussed above, it is reasonable to speculate that the elevated Gal-1 in pregnancy may contribute to reduced uterine arterial myogenic tone by suppressing Cav1.2 surface expression/activity [209,210]. Intriguingly, both circulating and placental Gal-1 is downregulated in early onset preeclampsia and pregnant women with fetal growth restriction [211,212]. The downregulation of Gal-1 probably plays a role in the increased myogenic tone in uterine and other vascular beds in preeclampsia due to upregulating Cav1.2 surface expression/activity [165,166,167,179].

4. Roles of Reactive Oxygen Species (ROS) in the Pathogenesis of Hypertensive Disorders

4.1. Overview of ROS

Reactive oxygen species (ROS) are oxygen-containing molecules naturally produced in cellular metabolism. ROS comprise free radicals such as superoxide anion (O2•−) and hydroxyl radical (OH) and nonradical molecule hydrogen peroxide (H2O2). O2•− is formed from the one-electron reduction of molecular oxygen (O2). It is a precursor to a cascade of other ROS. Its dismutation, either occurring spontaneously or being catalyzed by superoxide dismutases (SODs), produces H2O2. Through the Fenton reaction, H2O2 can be reduced to OH. Moreover, the reaction between O2•− and nitric oxide (NO) results in the formation of peroxynitrite (ONOO).
ROS are produced in different cellular compartments including mitochondria, endoplasmic reticulum (ER), lysosomes, peroxisomes, and plasma membrane [23,213,214]. Nicotinamide adenine dinucleotide phosphate oxidases (NOXs) and mitochondria are the major sources of ROS [215] (Figure 3). NOXs are a family of transmembrane proteins that catalyze the transfer of electron donated by NADPH to molecular oxygen to form O2•− [216]. NOXs comprise seven isoforms (NOX1-5 and dual oxidases (DUOX) 1 and 2). NOX1, NOX2, NOX4, and NOX5 are primary isoforms in vasculature [217]. During oxidative phosphorylation, ATP synthesis is coupled to the movement of electrons derived from NADH and FADH2 through Complexes I to IV in the mitochondrial electron transport chain (ETC). Electrons are received by O2 at Complex IV to produce H2O. Mitochondria consumes ~90% cellular O2 [218]. It is estimated that ~1–2% of O2 consumed by mitochondria reacted with leaked electrons, leading to the formation of O2•− [219]. Complexes I and III are the major sites in ETC to generate O2•− [220].
ROS can be beneficial and detrimental. They are highly reactive. O2•− and OH are short-lived and exert their actions in their immediate vicinity. In contrast, H2O2 is stable and can travel a long distance in cells and pass through the membrane. While O2•− and OH nonspecifically act on macromolecules such as proteins, DNA, and lipids, H2O2 can trigger reversible oxidation of cysteine residues in specific proteins including enzymes, protein kinases, transcription factors, and ion channels, leading to altered protein functions [221,222]. Thus, H2O2 is considered as an intracellular signaling molecule at physiological concentrations [221,222]. However, H2O2 causes irreversible damage to macromolecules and impairs their functions at supraphysiological concentrations.
Cellular ROSs are tightly regulated under physiological conditions. Physiological ROS concentrations are maintained by various antioxidant mechanisms. O2•− is converted into H2O2 by Cu, Zn-SOD (SOD1) in the cytosol/mitochondrial intermembrane space and Mn-SOD (SOD2) in the mitochondrial matrix. H2O2 is subsequently decomposed to H2O by catalase, glutathione peroxidases (GPXs) and peroxiredoxins (PRXs). ROS can also be scavenged by nonenzymatic antioxidants such as glutathione (GSH), vitamins C/E, uric acid, and melatonin [223]. Oxidative stress occurs when the ROS homeostasis is disrupted due to ROS overproduction, or antioxidant deficiency, or both. Thus, oxidative stress represents a state of an imbalance between ROS production and antioxidant defenses.

4.2. Oxidative Stress as a Hallmark in Hypertension-Related Disorders

Hypertension, diabetes, and preeclampsia exhibit a state of oxidative stress [21,22,23]. A variety of studies have demonstrated that levels of biomarkers of oxidative stress such as O2•−/H2O2, malondialdehyde, 8-isoprostaglandin F2α(8-iso-PGF2α), oxidized low-density lipoprotein (ox-LDL), and protein carbonyl are increased, whereas levels of antioxidants GSH (or GSH/GSSG ratio), vitamin C/E, and levels/activities of SOD, catalase, and GRX are decreased in the circulation of hypertensive [224,225,226,227,228,229,230,231], diabetic [232,233,234,235,236,237,238,239], and preeclamptic [240,241,242,243,244,245,246,247,248] patients. In addition, urinary 8-iso-PGF2α excretion is also increased in these disorders [231,249,250]. Systemic oxidative stress has also been detected in animal models of hypertension [251,252,253,254,255,256,257,258], diabetes [259,260,261,262], and preeclampsia [263,264,265].
Notably, hypertension-related disorders also undergo vascular oxidative stress. Both O2•− and H2O2 are increased in vascular cells in experimental models of hypertension including SHR/SHRSP rats, DOCA rats, angiotensin II-induced mice, and transgenic mice overexpressing smooth muscle cell-specific NOX1 [266,267,268,269,270,271,272,273,274,275]. In addition, SHR and angiotensin II-infused mice display elevated vascular lipid peroxidation, 8-hydroxydeoxyguanosine, and nitrotyrosine [252,276]. Chronic administration of aldosterone increases O2•− production in mouse cerebral arteries and aortas [277,278]. Moreover, specific deletion of the mineralocorticoid receptor suppresses angiotensin II-induced vascular ROS [142,143]. Diabetic patients have elevated vascular O2•− [279]. Similarly, vascular O2•− and H2O2 are increased in db/db mice and GK rats [280,281,282]. The increased vascular ROS are simulated by exposing vascular smooth muscle cells and endothelial cells to high glucose [283,284,285,286,287]. O2•− overproduction and/or oxidative stress biomarkers nitrotyrosine and 4-hydroxnonenal (4-HNE) are observed in the placenta and vasculature of women with preeclampsia [288,289,290,291,292]. Placental and vascular oxidative stress are also detected in animal models of preeclampsia including catechol-O-methyltransferase deficient (COMT−/−) mice, endothelial nitric oxide synthase deficient (eNOS−/−) mice, RUPP rats, rodents/sheep experiencing gestational hypoxia, and rodents infused with preeclampsia-related bioactive factors [264,265,293,294,295,296,297,298,299,300,301,302,303]. Exposure to plasma from women with preeclampsia promotes O2•− production in uterine arteries and vascular cells [246,304]. Ex vivo hypoxia also induces heightened oxidative stress in uterine arteries [305].
ROS overproduction by NOXs and mitochondria are major sources leading to the heightened oxidative stress in vasculature of human hypertensive patients and experimental animal models of hypertension-related disorders, although suppressed activity and/or expression of SODs, catalase, GPX, TRX, and TRXR also contribute [276,286,288,289,290,293,306,307,308,309].

4.2.1. NOX-Derived ROS in Hypertension-Related Disorders

Several polymorphisms in the gene encoding p22phox in human are associated with hypertension by affecting enzymatic activity [310]. In SHR/SHRSP, the expression/activity of vascular NOXs 1, 2 and 4 are increased [270,272,311,312,313,314,315,316]. DOCA rats have higher vascular expression of NOX subunit p22phox and enzymatic activity [270,317]. NOX activity in cultured vascular smooth cells and in cultured rat mesangial cells is also stimulated by aldosterone, leading to increased ROS generation [318,319]. Chronic administration of aldosterone causes NOX2-depednet increase O2•− production in mouse cerebral arteries [277]. Dexamethasone upregulates Nox1 expression in rat aorta via activation of the glucocorticoid receptor [320]. In a mouse model of hyperadrenergic hypertension created by targeted ablation of the chromogranin a (Chga) gene, renal expression of NOX1/2 is increased [321]. Angiotensin II-infused rodents are associated with elevated expression of NOXs 1 and 2 as well as NOX subunits p22phox, p47phox, and Rac1 in vessels and increased NOX activity [268,274,322,323,324,325,326,327]. Angiotensin II apparently contributes to the upregulation of NOXs as it stimulates gp91phox (NOX2) and p22phox expression in vascular smooth muscle cells from human resistance arteries [328].
In diabetic patients, vascular protein abundance of NOX subunits p22phox, p67phox, and p47phox and enzymatic activity of NOXs are increased [279]. Moreover, O2•− production is reduced by non-selective NOX inhibitor diphenylene iodonium in vessels from diabetic patients [279]. The expression and activity of NOXs 1, 2 and 4 as well as p22phox/p47phox are increased in aorta and mesenteric arteries of STZ rodents [287,308,329,330,331,332,333]. Vascular expression/activity of NOXs 1, 2, and 4 as well as p22phox are also elevated in db/db mice [280,334,335]. Chronic high glucose exposure stimulates p22phox, p47phox and p67phox expression and NOX activity in cultured endothelial cells [285,286,336,337,338]. Similarly, high glucose also promotes NOXs 1 and 4 as well as p22phox expression in cultured vascular smooth muscle cells [287,308].
In human, the preeclamptic placenta exhibits elevated expression of NOX1, p22phox, p47phox, and p67phox [308,339]. Preeclampsia is also associated with increased vascular NOX2 expression [304,340]. The expression of NOX2 in ovine uterine arteries is also higher in high-altitude pregnancy [299]. Women with preeclampsia produce excessive angiotensin II type 1-receptor autoantibody (AT1-AA), soluble fms-like tyrosine kinase-1 (sFlt-1), ox-LDL, and tumor necrosis factor-α (TNF-α), leading to high concentrations of them in the circulation [244,341,342,343,344]. In rat models of preeclampsia induced by AT1-AA or sFlt-1 infusion, NOX activity is found to be increased in the placenta, kidney, and aorta [294,345,346,347,348]. sFlt-1 promotes the expression of NOXs 1, 2, and 4 in the placenta and kidney [348]. Circulating activin is also increased in preeclampsia and activin infusion into pregnant mice enhances NOX2 expression in aorta [340,349]. In cultured human trophoblasts, AT1-AA promotes the expression of p22phox, p47phox and p67phox and increases O2•− generation [350]. In cultured human uterine microvascular endothelial cells and human umbilical vein endothelial cells (HUVECs), preeclamptic plasma/serum increases the expression of NOXs 2 and 4 and NOX activity, which is simulated by sFlt-1, activin, and ox-LDL [292,340,345,351]. The expression of p47phox and p67phox rises and production of O2•− increases following AT1-AA treatment in cultured human vascular smooth muscle cells [350]. TNF-α boosts p22phox express and O2•− generation in rat aortic smooth muscle cells [352]. Ox-LDL stimulates vascular ROS generation via both upregulating its receptor lectin-like ox-LDL receptor-1 (LOX-1) and stimulating NOX activity [353,354]. Vascular O2•− production is increased in RUPP rats in part due to LOX-1 upregulation [296].

4.2.2. Mitochondria-Derived ROS in Hypertension-Related Disorders

Mitochondria become dysfunctional in hypertension-related disorders. SIRT3, a histone deacetylase, is an important regulator of mitochondrial redox state. It interacts with SOD2 in mitochondria and subsequently promotes SOD2 deacetylation, leading to enhanced enzymatic activity [355,356]. Essential hypertension in human is associated with increased mitochondrial oxidative stress in arterioles from mediastinal fat due to reduced SIRT3 and increased SOD2 acetylation [357]. Similar to human essential hypertension, SHR also exhibits overproduction of mitochondrial ROS in vessels [358]. Partial deletion of SOD2 in mitochondria (SOD2+/−) increases renal oxidative stress and blood pressure [359]. In angiotensin II-infused mice, the increased blood pressure is associated with increased SIRT3 S-glutathionylation and vascular SOD2 acetylation, reduced SOD2 activity and elevated vascular O2•− production [357,360]. Those alterations are diminished by SIRT3 overexpression [357]. Cyclophilin D is a regulatory subunit of the mitochondrial permeability transition pore and participates in regulating mitochondrial function [361]. Evidently, it is an important mediator of angiotensin II-induced hypertension as its depletion diminishes both mitochondrial O2•− generation in aorta and hypertension [362]. Mitochondrial O2•− generation in DOCA rat mesenteric arteries is also increased [255]. In cultured HUVECs, excess aldosterone suppresses SOD2 expression and increases mitochondrial ROS production [363]. Interestingly, mitochondrial ROS is also regulated by a phenomenon termed ROS-induced ROS release [364]. In cultured human aortic endothelial cells, silence of NOX2 with siRNA reduces angiotensin II-induced mitochondrial O2•− generation [365]. Likewise, NOX2 deletion by gp91phox knockout also reduces angiotensin II-induced mitochondrial O2•− generation in aorta and hypertension [365].
Elevation of mitochondrial O2•− is detected in subcutaneous arterioles in type 2 diabetic patients [366]. Likewise, mitochondrial H2O2 is increased in primary human saphenous vein endothelial cells from type 2 diabetic subjects [367]. STZ mice/rats, HFD mice, ZDF rats, and GK rats display increased mitochondrial ROS in vascular smooth muscle cells and endothelial cells [331,368,369,370,371,372]. Hyperglycemia is apparently a causative factor leading to heightened mitochondrial ROS in vascular cells. Exposure to high concentrations of glucose stimulates mitochondrial ROS production in a variety of endothelial cells, promoting mitochondrial damage by producing 8-hydroxydeoxyguanosine and nitrotyrosine [373,374,375,376,377,378,379,380,381].
Preeclampsia is associated with reduced mitochondrial content and decreased oxidative phosphorylation in the placenta [382,383]. Preeclamptic placenta also exhibits mitochondrial oxidative stress as evidenced by increased malondialdehyde and 4-HNE-modified proteins in mitochondria [384,385]. Mitochondrial SOD2 activity is reduced in preeclamptic placenta [309,386]. Mitochondrial ROS is also elevated in the kidney and placenta of RUPP rats [387]. Mitochondrial ROS in the placenta is raised by infusing/injecting with either of AT1-AA, sFlt-1, TNFα and ad-CMV-hypoxia-inducible factor (HIF) into pregnant rodents [302,388,389,390]. In cultured endothelial cells, serum from women with preeclampsia suppresses mitochondrial respiration and boosts mitochondrial ROS generation [391,392,393]. In a way like preeclamptic serum, sFlt-1 also promotes mitochondrial oxidative stress in endothelial cells [393]. Serum from RUPP rats similarly stimulates mitochondrial ROS production in HUVECs, which is ablated by AT1-AA inhibition [387,388]. Hypoxic treatment reduces mitochondrial respiration in cultured trophoblast-like JEG3 cells and increases mitochondrial ROS in human placenta [394,395]. In vitro hypoxia also suppresses mitochondrial respiration and increases mitochondrial ROS in ovine uterine artery smooth muscle cells [396].

4.3. A Causative Role of ROS in in Animal Models of Hypertension-Related Disorders

Oxidative stress is believed to be linked to pathogenesis and progression of a myriad of human diseases including hypertension-related disorders [21,23,24,397,398,399,400,401]. Numerous studies using animal models reveal a causative role of ROS in the pathogenesis of hypertension, diabetes, and preeclampsia.
Treatment with the SOD mimetic tempol reduces vascular ROS and lower blood pressure in animal models of essential and secondary hypertension [253,257,269,317,402,403,404], of diabetes [405,406,407,408], and of preeclampsia [293,294,346,409]. SOD1 deletion promotes development of hypertension, whereas the delivery of liposome-encapsulated SOD diminishes Ang II-induced hypertension [267,410].
NOX inhibitors apocynin and diphenyleneiodonium have been shown to decrease vascular ROS, improve vascular function, and lower blood pressure in animal models of hypertension-related disorders [272,275,312,317,324,340,411,412,413,414,415]. Angiotensin II-induced vascular ROS and hypertension is enhanced in transgenic mice overexpressing NOX1 in smooth muscle cells and is suppressed in NOX1 deficient mice [274,325,416]. Similarly, angiotensin II-induced vascular O2•− and hypertension are blunted in p47phox−/− mice [271]. Deletion of p47phox prevents diabetes-induced vascular dysfunction in mice [332,417]. Similar findings are also observed in db/db mice with siRNA-induced p22phox knockdown [334].
Administration of mitochondria-targeted antioxidants have extensively explored in animal models of hypertension and preeclampsia. MitoQ decreases blood pressure in SPSHR [418], whereas MitoEbselen reduces vascular ROS and prevents angiotensin II-induced hypertension in mice [360]. MitoQ and MitoTempol also reduce placental oxidative stress, lowers blood pressure, and improves fetal growth in RUPP rats and in pregnant rats experiencing gestational hypoxia [301,387,389]. Scavenging of mitochondrial isolevuglandins formed by peroxidation of arachidonic acid with Mito2HOBA and targeting cyclophilin D, a component of mitochondrial permeability transition pore, with its specific inhibitor sanglifehrin A reduce vascular mitochondrial ROS, improve vascular function, and attenuate hypertension in Ang II-infused mice [362,419]. These beneficial responses are replicated by overexpression of mitochondria-targeted SOD2 and catalase [362].

5. Contribution of Dysfunctional Cav1.2 Conferred by ROS to Increased Vascular Tone in Hypertension-Related Disorders

5.1. ROS and Cav1.2 Function/Expression

There is ample evidence that Cav1.2 is regulated by cellular redox state. The α1c subunit contains 48 cysteines with some of them being sensitive to redox modulation [420]. Redox modulation of sulfhydryl groups of cysteine residues could alter the structure and function of proteins. Site-directed mutagenesis has identified several cysteine residues including C543, C1789, C1790, and C1810 that subject to redox regulation [421,422]. However, electrophysiological studies reveal conflicting effects of ROS on Cav1.2. Studies from Amberg’s group demonstrate that bath application of H2O2 and ROS produced by xanthine oxidase/hypoxanthine stimulates Cav1.2 activity in rat cerebral arterial smooth muscle cells [423,424]. Both mitochondria- and NOX-derived ROS could exert stimulatory effects on Cav1.2. Antimycin enhances Cav1.2 activity by promoting mitochondrial H2O2 generation. In addition, angiotensin II augments Cav1.2 activity by promoting mitochondria- and NOX-mediated H2O2 production [423,424,425]. Rotenone also increases Cav1.2-mediated currents by stimulating mitochondrial O2•− production in A7r5 aortic smooth muscle cells [426]. In contrast, sulfhydryl-oxidizing agents 2,2′-dithiodipyridine (DTDP) and thimerosal inhibit heterologously expressed α1c subunit of rabbit smooth muscle Cav1.2 [427]. In rat tail arterial smooth muscle cells, 2,5-di-t-butyl-1,4-benzohydroquinone inhibits Cav1.2 via O2•− generation [428]. H2O2 is found to inhibit Cav1.2 in A7r5 aortic smooth muscle cells [426]. The disunity in electrophysiological findings suggests complexity of ROS actions, possibly due to concentrations of ROS used, the expression systems, ROS compartmentation and absence/presence of channel subunits/partners.
The redox modulation of Cav1.2 has immense impacts on vascular function. H2O2 is reported to trigger an increase in [Ca2+]i and vasoconstriction in rat mesenteric arteries, rat coronary arteries, rat aorta, and canine cerebral arteries [429,430,431,432,433,434,435]. H2O2-induced increase in [Ca2+]i and contractile response are inhibited by the removal of extracellular Ca2+ and by Cav1.2 inhibitors nisoldipine, nifedipine, diltiazem, and verapamil, suggesting that H2O2 activates Cav1.2 in vascular smooth muscle cells [429,430,431,434]. Likewise, an ONOO-induced increase in [Ca2+]i is inhibited by nifedipine in mesenteric arteriolar smooth muscle cells [436]. Thrombin stimulates anincrease of [Ca2+]i via activating Cav1.2 in rat aortic smooth muscle cells, which is absent in NOX1 null cells [437]. As expected, vessels from SHR exhibits greater increase in [Ca2+]i and vasoconstriction compared to WKY [431,432,433,435]. It should be noted that H2O2 also functions as an EDHF in other vascular beds including mouse and human mesenteric arteries, mouse aorta, and in porcine coronary microvessels and causes vasorelaxation [438,439,440,441]. Depolarization with high concentrations of KCl causes vasoconstriction by activating Cav1.2 [442]. Paraquat and LY83583 enhance KCl-induced increase in [Ca2+]i and vasoconstriction by increasing O2•− production in rat renal afferent arterioles, mesenteric arteries and intrapulmonary arteries [443,444]. Sphingosylphosphorylcholine also potentiates the contractile response to KCl in a NOX-dependent mechanism in rat mesenteric and intrapulmonary arteries [443].
NO is also a reactive, gaseous signaling molecule. S-nitrosylation, the covalent attachment of NO to a cysteine thiol in a given protein, is also an important redox signaling [445]. S-nitrosylation of Cys 1180 and/or Cys1280 in Cav1.2 reduces surface expression of the channel and channel activity [446]. In addition, S-nitrosylation of Cav1.2 also reduces channel open probability [447]. Excess ROS leads to low NO bioavailability in hypertension [448]. Intriguingly, Cav1.2 S-nitrosylation levels are reduced in aorta from SHR rats and in pulmonary arteries from patients with pulmonary hypertension [446]. Not surprisingly, increasing vascular s-nitrosylation levels of Cav1.2 by injecting S-nitrosocysteine reduces blood pressure in SHR rats [446].
ROS generated in different cellular compartments form microdomains and act on their targets in a spatiotemporal manner [449] (Figure 4). The redox regulation of Cav1.2 requires adjacent location of mitochondria and NOXs to Cav1.2 on the plasma membrane in vascular smooth muscle cells [423,425]. Thus, it is expected that microdomains with high concentrations of ROS exist in subcellular regions next to Cav1.2. Interestingly, redox modulation of Cav1.2 can occur directly or indirectly. In addition to directly targeting Cav1.2, ROS could oxidize partners of Cav1.2 to indirectly alter Cav1.2 activity. As aforementioned, PKC and c-Src form a partnership with Cav1.2 to regulate the activity of the channel. PKC contains six cysteine residues in each of the two pairs of zinc fingers within the regulatory domain and the oxidation of them activates PKCs [450,451]. H2O2 has been shown to activate PKCs in vascular smooth muscle cells [452,453]. SH2 domain of c-Src also contains cysteine residues and oxidation of these residues promotes the formation of disulfide bridges that subsequently override the autoinhibitory effect of Tyr527, leading to enzymatic activation [454]. Vascular c-Src is activated by NOX4- and mitochondria-derived ROS [287,455]. ROS produced by both NOXs and mitochondria are found to activate adjoining PKCs, which in turn enhances Cav1.2 activity in vascular smooth muscle cells [423,425,426]. Not surprisingly, H2O2-induced vasocontraction is reduced by PKC/c-Src inhibition and by Cav1.2 blockade [430,434,435,456,457]. Similarly, the potentiation of KCl-induced vasoconstriction of rat mesenteric and intrapulmonary arteries by NOX-derived ROS is also sensitive to the inhibition of PKC/c-Src [443]. Calmodulin is a key Ca2+ sensor for both Cav1.2 inactivation and facilitation [458]. Modulation of calmodulin by H2O2 also augments cardiac Cav1.2 activity [459].
Remarkably, ROS also regulate gene expression [460,461]. There is a concurring decrease in both vascular ROS and Cav1.2 expression in aortas of aged mice lacking the mineralocorticoid receptor in smooth muscle cells [142]. The protein expression of the α1C subunit is increased by angiotensin II in cultured rat mesenteric arteries [462,463]. H2O2 mimics, while NOX inhibition by apocynin, diphenyleneiodonium and gp91ds-tat as well as catalase annuls angiotensin II-mediated upregulation of the α1C subunit [463]. In a cardiac muscle cell line HL-1 cells, angiotensin II upregulates the α1C subunit at both mRNA and protein levels through the NOX-PKC pathway-mediated cAMP response element binding protein (CREB) phosphorylation [464]. It is expected that this mechanism could also play a role in secondary hypertension as hypertension due to Cushing’s syndrome, hyperaldosteronism, and renovascular disease involves activation of the renin-angiotensin system [231,256,465,466,467,468,469]. In rat cerebral arteries, endothelin 1 promotes mitochondrial ROS generation which subsequently stimulates the expression of the α1C subunit by activating nuclear factor kappa B (NF-κB) [470]. Consequently, the upregulation of Cav1.2 induced by endothelin 1 increases both myogenic tone and depolarization-induced vasoconstriction.

5.2. ROS and Myogenic Tone

Exogenous ROSs have been shown to alter myogenic tone. In pressurized mouse tail arterioles and rat gracilis skeletal muscle arterioles, H2O2 causes vasoconstriction [471,472]. The exposure of rat cerebral arteries to the ROS-generating system xanthine oxidase plus hypoxanthine also increases myogenic tone [423]. Likewise, O2•− generated by paraquat enhances myogenic tone in mouse afferent arterioles [473]. OH generated by the Fenton reaction from H2O2 and the iron redox chelate Fe3+/nitrilotriacetate (FeNTA) elevates myogenic tone in denuded rat ophthalmic arteries [474].
Numerous studies reveal that endogenous ROS plays a key role in intralumenal pressure-induced myogenic tone. Myogenic tone in mesenteric arteries from SOD1 knockout mice is increased [475]. ROS derived from both NOXs and mitochondrial contribute to myogenic tone development. Pressurization of rat femoral arterial branches increases O2•− via activating NOXs [476]. Similarly, an increase in intralumenal pressure promotes ROS in vascular smooth muscle cells and myogenic constriction simultaneously in mouse tail arterioles, which are inhibited by diphenyleneiodonium/catalase and are absent in arterioles from p47phox−/− mice or transgenic mice expressing a dominant-negative mutant of human rac1 (rac-DN) [471]. Likewise, pressure-dependent myogenic tone enhanced in mouse afferent arterioles is repressed in vessels from p47phox−/− mice and by NOX inhibitors apocynin and diphenyleneiodonium [477,478]. The regulation of myogenic tone by ROS is tissue-, ROS type- and ROS source-dependent. H2O2, but not O2•−, contributes to myogenic constrictor response in rat tail arterioles as the vasoconstriction is diminished by catalase, but not by manganese(III) tetrakis(1-methyl-4-pyridyl)porphyrin (MnTMPyP), a cell-permeant mimic of SOD [471]. However, myogenic tone is reduced by pegylated SOD, but not by pegylated catalase in mouse afferent arterioles [477]. Selective inhibition of both NOX1 and NOX2 with 2-acetylphenothiazine (2-APT) and VAS2870, respectively, suppressed basal myogenic tone in rat cremaster skeletal muscle arterioles [479], whereas myogenic tone is inhibited by NOX1 inhibitor ML171 but not by NOX2 inhibitor gp91-dstat in rat cerebral arteries [480]. The concomitant increases in superoxide/H2O2 and myogenic constriction induced by pressure is also reduced by MitoQ in rat cerebral arteries [481].
As expected, both O2•− and myogenic tone are increased in vessels from animal models of hypertension-related disorders including SHR cerebral arteries/afferent arterioles, OZR gracilis skeletal muscle arterioles/cerebral arteries, and db/db mouse mesenteric arteries [131,138,482,483,484,485]. The increased myogenic constriction of mesenteric arteries of db/db mice is reduced by supplement of GSH or the Nrf2-activator sulforaphane [485]. A reduction in myogenic tone is also achieved by scavenging ROS in SHR afferent arterioles, db/db gracilis skeletal muscle arterioles, and OZR cerebral arteries/gracilis skeletal muscle arterioles [131,281,482,483,484]. Aging is associated with increased prevalence of hypertension and increases in intralumenal pressure promote mitochondrial oxidative stress in aged mouse cerebral arteries [486]. The exacerbated myogenic tone of uterine arteries from high-altitude pregnant sheep concurs with increased expression of NOX2 and mitochondrial ROS, and pressure-induced myogenic tone in uterine arteries from high-altitude pregnant sheep is diminished by apocynin/mitoQ, whereas it is enhanced in uterine arteries from low-altitude pregnant sheep by mitochondrial respiratory chain inhibitors rotenone and antimycin [299,396].
Numerous studies demonstrate that PKC/c-Src activation contributes to myogenic tone development in small arteries/arterioles [136,487,488,489,490,491]. The expression/activity of vascular PKC/c-Src is increased in SHR and STZ rats as well as in uterine arteries of high-altitude pregnant sheep [179,435,492,493]. PKC/c-Src activation has a greater contribution to myogenic tone in SHR cerebral arteries, STZ rat gracilis skeletal muscle arterioles, and high-altitude sheep uterine arteries [136,159,179]. Notably, the increased myogenic constriction by ROS in rat cerebral arteries is mediated by ROS-mediated PKC/c-Src activation [423,473,494]. Moreover, PKC/c-Src activation in turn stimulates Cav1.2 activity to promote myogenic constriction [86,112,473]. Nevertheless, both PKCs and c-Src can also contribute to myogenic tone development by increasing the sensitization of the contractile apparatus to Ca2+ [112,136].

6. Conclusions/Perspective

Evidently, Ca2+ influx through Cav1.2 is fundamental to vasoconstriction and myogenic tone of small arteries and arterioles. Antihypertensive drugs including Cav1.2 blockers have been successfully used to the management of hypertension-related disorders [495,496,497,498]. Oxidative stress is a hallmark of hypertension-related disorders. However, it is still an open question whether oxidative stress is a cause or consequence of hypertension. The data reviewed here highlights critical roles of vascular oxidative stress in the pathophysiology of hypertension-related disorders. Vascular Cav1.2 is targeted by excessive ROS directly and indirectly leading to exaggerated channel expression/activity. The dysfunction of Cav1.2 ultimately results in increased myogenic tone and elevated blood pressure. In preclinical studies, ROS-induced myogenic tone is diminished by Cav1.2 blocker nifedipine [473]. Moreover, apocynin is as effective as nifedipine in lowering blood pressure in an animal model of hypertension [411]. Regulation of Cav1.2 by ROS primarily occurs in ROS microdomains [423,425]. In this context, selectively combating compartmentalized ROS could also be a viable therapeutic approach in the treatment of hypertension. Animal studies have provided compelling evidence supporting a causative role of ROS in the pathogenesis of hypertension-related disorders. However, clinical trials on humans provide less convincing evidence [23,499]. Although animal studies have significantly contributed to our understanding mechanisms underlying hypertension-related disorders, further work is needed to translate findings from animal research to clinical benefit.

Author Contributions

Conceptualization, X.-Q.H. and L.Z.; writing—original draft preparation, X.-Q.H. and L.Z.; writing—review and editing, X.-Q.H. and L.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Institutes of Health Grants HD083132 (L.Z.) and HL149608 (L.Z.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Berridge, M.J.; Lipp, P.; Bootman, M.D. The versatility and universality of calcium signalling. Nat. Rev. Mol. Cell Biol. 2000, 1, 11–21. [Google Scholar] [CrossRef] [PubMed]
  2. Woll, K.A.; Van Petegem, F. Calcium-release channels: Structure and function of IP3 receptors and ryanodine receptors. Physiol. Rev. 2022, 102, 209–268. [Google Scholar] [CrossRef] [PubMed]
  3. Ghosh, D.; Syed, A.U.; Prada, M.P.; Nystoriak, M.A.; Santana, L.F.; Nieves-Cintron, M.; Navedo, M.F. Calcium Channels in Vascular Smooth Muscle. Adv. Pharmacol. 2017, 78, 49–87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Ottolini, M.; Hong, K.; Sonkusare, S.K. Calcium signals that determine vascular resistance. Wiley Interdiscip Rev. Syst. Biol. Med. 2019, 11, e1448. [Google Scholar] [CrossRef] [PubMed]
  5. Whelton, P.K.; Carey, R.M.; Aronow, W.S.; Casey, D.E., Jr.; Collins, K.J.; Dennison Himmelfarb, C.; DePalma, S.M.; Gidding, S.; Jamerson, K.A.; Jones, D.W.; et al. 2017 ACC/AHA/AAPA/ABC/ACPM/AGS/APhA/ASH/ASPC/NMA/PCNA Guideline for the Prevention, Detection, Evaluation, and Management of High Blood Pressure in Adults: Executive Summary: A Report of the American College of Cardiology/American Heart Association Task Force on Clinical Practice Guidelines. Hypertension 2018, 71, 1269–1324. [Google Scholar] [CrossRef]
  6. Brown, M.A.; Magee, L.A.; Kenny, L.C.; Karumanchi, S.A.; McCarthy, F.P.; Saito, S.; Hall, D.R.; Warren, C.E.; Adoyi, G.; Ishaku, S.; et al. Hypertensive Disorders of Pregnancy: ISSHP Classification, Diagnosis, and Management Recommendations for International Practice. Hypertension 2018, 72, 24–43. [Google Scholar] [CrossRef] [Green Version]
  7. Justin, J.; Fayol, A.; Bruno, R.M.; Khettab, H.; Boutouyrie, P. International Guidelines for Hypertension: Resemblance, Divergence and Inconsistencies. J. Clin. Med. 2022, 11, 1975. [Google Scholar] [CrossRef]
  8. de Boer, I.H.; Bangalore, S.; Benetos, A.; Davis, A.M.; Michos, E.D.; Muntner, P.; Rossing, P.; Zoungas, S.; Bakris, G. Diabetes and Hypertension: A Position Statement by the American Diabetes Association. Diabetes Care 2017, 40, 1273–1284. [Google Scholar] [CrossRef] [Green Version]
  9. Tsimihodimos, V.; Gonzalez-Villalpando, C.; Meigs, J.B.; Ferrannini, E. Hypertension and Diabetes Mellitus: Coprediction and Time Trajectories. Hypertension 2018, 71, 422–428. [Google Scholar] [CrossRef]
  10. Shin, D.; Bohra, C.; Kongpakpaisarn, K. Impact of the Discordance Between the American College of Cardiology/American Heart Association and American Diabetes Association Recommendations on Hypertension in Patients With Diabetes Mellitus in the United States. Hypertension 2018, 72, 256–259. [Google Scholar] [CrossRef]
  11. Whelton, P.K. The elusiveness of population-wide high blood pressure control. Annu. Rev. Public Health 2015, 36, 109–130. [Google Scholar] [CrossRef]
  12. Mills, K.T.; Bundy, J.D.; Kelly, T.N.; Reed, J.E.; Kearney, P.M.; Reynolds, K.; Chen, J.; He, J. Global Disparities of Hypertension Prevalence and Control: A Systematic Analysis of Population-Based Studies From 90 Countries. Circulation 2016, 134, 441–450. [Google Scholar] [CrossRef]
  13. Mills, K.T.; Stefanescu, A.; He, J. The global epidemiology of hypertension. Nat. Rev. Nephrol. 2020, 16, 223–237. [Google Scholar] [CrossRef]
  14. Rana, S.; Lemoine, E.; Granger, J.P.; Karumanchi, S.A. Preeclampsia: Pathophysiology, Challenges, and Perspectives. Circ. Res. 2019, 124, 1094–1112. [Google Scholar] [CrossRef]
  15. Chobanian, A.V.; Bakris, G.L.; Black, H.R.; Cushman, W.C.; Green, L.A.; Izzo, J.L., Jr.; Jones, D.W.; Materson, B.J.; Oparil, S.; Wright, J.T., Jr.; et al. Seventh report of the Joint National Committee on Prevention, Detection, Evaluation, and Treatment of High Blood Pressure. Hypertension 2003, 42, 1206–1252. [Google Scholar] [CrossRef] [Green Version]
  16. Rimoldi, S.F.; Scherrer, U.; Messerli, F.H. Secondary arterial hypertension: When, who, and how to screen? Eur. Heart J. 2014, 35, 1245–1254. [Google Scholar] [CrossRef] [Green Version]
  17. Chatterjee, S.; Khunti, K.; Davies, M.J. Type 2 diabetes. Lancet 2017, 389, 2239–2251. [Google Scholar] [CrossRef]
  18. Sakran, N.; Graham, Y.; Pintar, T.; Yang, W.; Kassir, R.; Willigendael, E.M.; Singhal, R.; Kooreman, Z.E.; Ramnarain, D.; Mahawar, K.; et al. The many faces of diabetes. Is there a need for re-classification? A narrative review. BMC Endocr. Disord. 2022, 22, 9. [Google Scholar] [CrossRef]
  19. Fuchs, F.D.; Whelton, P.K. High Blood Pressure and Cardiovascular Disease. Hypertension 2020, 75, 285–292. [Google Scholar] [CrossRef]
  20. Michael, S.K.; Surks, H.K.; Wang, Y.; Zhu, Y.; Blanton, R.; Jamnongjit, M.; Aronovitz, M.; Baur, W.; Ohtani, K.; Wilkerson, M.K.; et al. High blood pressure arising from a defect in vascular function. Proc. Natl. Acad. Sci. USA 2008, 105, 6702–6707. [Google Scholar] [CrossRef]
  21. Raijmakers, M.T.; Dechend, R.; Poston, L. Oxidative stress and preeclampsia: Rationale for antioxidant clinical trials. Hypertension 2004, 44, 374–380. [Google Scholar] [CrossRef] [PubMed]
  22. Sena, C.M.; Leandro, A.; Azul, L.; Seica, R.; Perry, G. Vascular Oxidative Stress: Impact and Therapeutic Approaches. Front. Physiol. 2018, 9, 1668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Griendling, K.K.; Camargo, L.L.; Rios, F.J.; Alves-Lopes, R.; Montezano, A.C.; Touyz, R.M. Oxidative Stress and Hypertension. Circ. Res. 2021, 128, 993–1020. [Google Scholar] [CrossRef] [PubMed]
  24. Giacco, F.; Brownlee, M. Oxidative stress and diabetic complications. Circ. Res. 2010, 107, 1058–1070. [Google Scholar] [CrossRef] [Green Version]
  25. Montezano, A.C.; Dulak-Lis, M.; Tsiropoulou, S.; Harvey, A.; Briones, A.M.; Touyz, R.M. Oxidative stress and human hypertension: Vascular mechanisms, biomarkers, and novel therapies. Can. J. Cardiol. 2015, 31, 631–641. [Google Scholar] [CrossRef] [PubMed]
  26. Touyz, R.M.; Alves-Lopes, R.; Rios, F.J.; Camargo, L.L.; Anagnostopoulou, A.; Arner, A.; Montezano, A.C. Vascular smooth muscle contraction in hypertension. Cardiovasc. Res. 2018, 114, 529–539. [Google Scholar] [CrossRef] [Green Version]
  27. Dolphin, A.C. Voltage-gated calcium channels and their auxiliary subunits: Physiology and pathophysiology and pharmacology. J. Physiol. 2016, 594, 5369–5390. [Google Scholar] [CrossRef] [Green Version]
  28. Catterall, W.A. Voltage-gated calcium channels. Cold Spring Harb. Perspect. Biol. 2011, 3, a003947. [Google Scholar] [CrossRef]
  29. Kamp, T.J.; Hell, J.W. Regulation of cardiac L-type calcium channels by protein kinase A and protein kinase C. Circ. Res. 2000, 87, 1095–1102. [Google Scholar] [CrossRef] [Green Version]
  30. Kobrinsky, E.; Tiwari, S.; Maltsev, V.A.; Harry, J.B.; Lakatta, E.; Abernethy, D.R.; Soldatov, N.M. Differential role of the alpha1C subunit tails in regulation of the Cav1.2 channel by membrane potential, beta subunits, and Ca2+ ions. J. Biol. Chem. 2005, 280, 12474–12485. [Google Scholar] [CrossRef]
  31. Campiglio, M.; Flucher, B.E. The role of auxiliary subunits for the functional diversity of voltage-gated calcium channels. J. Cell. Physiol. 2015, 230, 2019–2031. [Google Scholar] [CrossRef]
  32. Catterall, W.A.; Perez-Reyes, E.; Snutch, T.P.; Striessnig, J. International Union of Pharmacology. XLVIII. Nomenclature and structure-function relationships of voltage-gated calcium channels. Pharmacol. Rev. 2005, 57, 411–425. [Google Scholar] [CrossRef]
  33. Mikami, A.; Imoto, K.; Tanabe, T.; Niidome, T.; Mori, Y.; Takeshima, H.; Narumiya, S.; Numa, S. Primary structure and functional expression of the cardiac dihydropyridine-sensitive calcium channel. Nature 1989, 340, 230–233. [Google Scholar] [CrossRef]
  34. Biel, M.; Ruth, P.; Bosse, E.; Hullin, R.; Stuhmer, W.; Flockerzi, V.; Hofmann, F. Primary structure and functional expression of a high voltage activated calcium channel from rabbit lung. FEBS Lett. 1990, 269, 409–412. [Google Scholar] [CrossRef] [Green Version]
  35. Murakami, M.; Yamamura, H.; Suzuki, T.; Kang, M.G.; Ohya, S.; Murakami, A.; Miyoshi, I.; Sasano, H.; Muraki, K.; Hano, T.; et al. Modified cardiovascular L-type channels in mice lacking the voltage-dependent Ca2+ channel beta3 subunit. J. Biol. Chem. 2003, 278, 43261–43267. [Google Scholar] [CrossRef] [Green Version]
  36. Bannister, J.P.; Adebiyi, A.; Zhao, G.; Narayanan, D.; Thomas, C.M.; Feng, J.Y.; Jaggar, J.H. Smooth muscle cell alpha2delta-1 subunits are essential for vasoregulation by Cav1.2 channels. Circ. Res. 2009, 105, 948–955. [Google Scholar] [CrossRef] [Green Version]
  37. Cox, R.H.; Fromme, S. Expression of Calcium Channel Subunit Variants in Small Mesenteric Arteries of WKY and SHR. Am. J. Hypertens. 2015, 28, 1229–1239. [Google Scholar] [CrossRef] [Green Version]
  38. Kharade, S.V.; Sonkusare, S.K.; Srivastava, A.K.; Thakali, K.M.; Fletcher, T.W.; Rhee, S.W.; Rusch, N.J. The beta3 subunit contributes to vascular calcium channel upregulation and hypertension in angiotensin II-infused C57BL/6 mice. Hypertension 2013, 61, 137–142. [Google Scholar] [CrossRef] [Green Version]
  39. Chen, Y.H.; Li, M.H.; Zhang, Y.; He, L.L.; Yamada, Y.; Fitzmaurice, A.; Shen, Y.; Zhang, H.; Tong, L.; Yang, J. Structural basis of the alpha1-beta subunit interaction of voltage-gated Ca2+ channels. Nature 2004, 429, 675–680. [Google Scholar] [CrossRef]
  40. Jay, S.D.; Sharp, A.H.; Kahl, S.D.; Vedvick, T.S.; Harpold, M.M.; Campbell, K.P. Structural characterization of the dihydropyridine-sensitive calcium channel alpha 2-subunit and the associated delta peptides. J. Biol. Chem. 1991, 266, 3287–3293. [Google Scholar] [CrossRef]
  41. Cheng, X.; Pachuau, J.; Blaskova, E.; Asuncion-Chin, M.; Liu, J.; Dopico, A.M.; Jaggar, J.H. Alternative splicing of Cav1.2 channel exons in smooth muscle cells of resistance-size arteries generates currents with unique electrophysiological properties. Am. J. Physiol. Heart Circ. Physiol. 2009, 297, H680–H688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Cheng, X.; Liu, J.; Asuncion-Chin, M.; Blaskova, E.; Bannister, J.P.; Dopico, A.M.; Jaggar, J.H. A novel Cav1.2 N terminus expressed in smooth muscle cells of resistance size arteries modifies channel regulation by auxiliary subunits. J. Biol. Chem. 2007, 282, 29211–29221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Hofmann, F.; Flockerzi, V.; Kahl, S.; Wegener, J.W. L-type Cav1.2 calcium channels: From in vitro findings to in vivo function. Physiol. Rev. 2014, 94, 303–326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Zamponi, G.W.; Striessnig, J.; Koschak, A.; Dolphin, A.C. The Physiology, Pathology, and Pharmacology of Voltage-Gated Calcium Channels and Their Future Therapeutic Potential. Pharmacol. Rev. 2015, 67, 821–870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Hu, Z.; Liang, M.C.; Soong, T.W. Alternative Splicing of L-type Cav1.2 Calcium Channels: Implications in Cardiovascular Diseases. Genes 2017, 8, 344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Puri, T.S.; Gerhardstein, B.L.; Zhao, X.L.; Ladner, M.B.; Hosey, M.M. Differential effects of subunit interactions on protein kinase A- and C-mediated phosphorylation of L-type calcium channels. Biochemistry 1997, 36, 9605–9615. [Google Scholar] [CrossRef] [PubMed]
  47. Gerhardstein, B.L.; Puri, T.S.; Chien, A.J.; Hosey, M.M. Identification of the sites phosphorylated by cyclic AMP-dependent protein kinase on the beta 2 subunit of L-type voltage-dependent calcium channels. Biochemistry 1999, 38, 10361–10370. [Google Scholar] [CrossRef]
  48. Bunemann, M.; Gerhardstein, B.L.; Gao, T.; Hosey, M.M. Functional regulation of L-type calcium channels via protein kinase A-mediated phosphorylation of the beta(2) subunit. J. Biol. Chem. 1999, 274, 33851–33854. [Google Scholar] [CrossRef] [Green Version]
  49. Yang, L.; Liu, G.; Zakharov, S.I.; Bellinger, A.M.; Mongillo, M.; Marx, S.O. Protein kinase G phosphorylates Cav1.2 alpha1c and beta2 subunits. Circ. Res. 2007, 101, 465–474. [Google Scholar] [CrossRef] [Green Version]
  50. Li, Y.; Yang, H.; He, T.; Zhang, L.; Liu, C. Post-Translational Modification of Cav1.2 and its Role in Neurodegenerative Diseases. Front. Pharmacol. 2021, 12, 775087. [Google Scholar] [CrossRef]
  51. Ishikawa, T.; Hume, J.R.; Keef, K.D. Regulation of Ca2+ channels by cAMP and cGMP in vascular smooth muscle cells. Circ. Res. 1993, 73, 1128–1137. [Google Scholar] [CrossRef] [Green Version]
  52. Liu, H.; Xiong, Z.; Sperelakis, N. Cyclic nucleotides regulate the activity of L-type calcium channels in smooth muscle cells from rat portal vein. J. Mol. Cell. Cardiol. 1997, 29, 1411–1421. [Google Scholar] [CrossRef]
  53. Ruiz-Velasco, V.; Zhong, J.; Hume, J.R.; Keef, K.D. Modulation of Ca2+ channels by cyclic nucleotide cross activation of opposing protein kinases in rabbit portal vein. Circ. Res. 1998, 82, 557–565. [Google Scholar] [CrossRef] [Green Version]
  54. Keef, K.D.; Hume, J.R.; Zhong, J. Regulation of cardiac and smooth muscle Ca2+ channels (Cav1.2a,b) by protein kinases. Am. J. Physiol. Cell Physiol. 2001, 281, C1743–C1756. [Google Scholar] [CrossRef]
  55. Fusi, F.; Manetti, F.; Durante, M.; Sgaragli, G.; Saponara, S. The vasodilator papaverine stimulates L-type Ca2+ current in rat tail artery myocytes via a PKA-dependent mechanism. Vasc. Pharmacol. 2016, 76, 53–61. [Google Scholar] [CrossRef]
  56. Nystoriak, M.A.; Nieves-Cintron, M.; Patriarchi, T.; Buonarati, O.R.; Prada, M.P.; Morotti, S.; Grandi, E.; Fernandes, J.D.; Forbush, K.; Hofmann, F.; et al. Ser1928 phosphorylation by PKA stimulates the L-type Ca2+ channel Cav1.2 and vasoconstriction during acute hyperglycemia and diabetes. Sci. Signal. 2017, 10, eaaf9647. [Google Scholar] [CrossRef] [Green Version]
  57. Fusi, F.; Mugnai, P.; Trezza, A.; Spiga, O.; Sgaragli, G. Fine tuning by protein kinases of Cav1.2 channel current in rat tail artery myocytes. Biochem. Pharmacol. 2020, 182, 114263. [Google Scholar] [CrossRef]
  58. Syed, A.U.; Reddy, G.R.; Ghosh, D.; Prada, M.P.; Nystoriak, M.A.; Morotti, S.; Grandi, E.; Sirish, P.; Chiamvimonvat, N.; Hell, J.W.; et al. Adenylyl cyclase 5-generated cAMP controls cerebral vascular reactivity during diabetic hyperglycemia. J. Clin. Investig. 2019, 129, 3140–3152. [Google Scholar] [CrossRef] [Green Version]
  59. Xiong, Z.; Sperelakis, N.; Fenoglio-Preiser, C. Regulation of L-type calcium channels by cyclic nucleotides and phosphorylation in smooth muscle cells from rabbit portal vein. J. Vasc. Res. 1994, 31, 271–279. [Google Scholar] [CrossRef]
  60. Taguchi, K.; Ueda, M.; Kubo, T. Effects of cAMP and cGMP on L-type calcium channel currents in rat mesenteric artery cells. Jpn. J. Pharmacol. 1997, 74, 179–186. [Google Scholar] [CrossRef]
  61. Quignard, J.F.; Frapier, J.M.; Harricane, M.C.; Albat, B.; Nargeot, J.; Richard, S. Voltage-gated calcium channel currents in human coronary myocytes. Regulation by cyclic GMP and nitric oxide. J. Clin. Investig. 1997, 99, 185–193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Sharma, N.; Bhattarai, J.P.; Hwang, P.H.; Han, S.K. Nitric oxide suppresses L-type calcium currents in basilar artery smooth muscle cells in rabbits. Neurol. Res. 2013, 35, 424–428. [Google Scholar] [CrossRef] [PubMed]
  63. Schuhmann, K.; Groschner, K. Protein kinase-C mediates dual modulation of L-type Ca2+ channels in human vascular smooth muscle. FEBS Lett. 1994, 341, 208–212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Callaghan, B.; Koh, S.D.; Keef, K.D. Muscarinic M2 receptor stimulation of Cav1.2b requires phosphatidylinositol 3-kinase, protein kinase C, and c-Src. Circ. Res. 2004, 94, 626–633. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Navedo, M.F.; Amberg, G.C.; Votaw, V.S.; Santana, L.F. Constitutively active L-type Ca2+ channels. Proc. Natl. Acad. Sci. USA 2005, 102, 11112–11117. [Google Scholar] [CrossRef] [Green Version]
  66. Cobine, C.A.; Callaghan, B.P.; Keef, K.D. Role of L-type calcium channels and PKC in active tone development in rabbit coronary artery. Am. J. Physiol. Heart Circ. Physiol. 2007, 292, H3079–H3088. [Google Scholar] [CrossRef]
  67. Nieves-Cintron, M.; Amberg, G.C.; Navedo, M.F.; Molkentin, J.D.; Santana, L.F. The control of Ca2+ influx and NFATc3 signaling in arterial smooth muscle during hypertension. Proc. Natl. Acad. Sci. USA 2008, 105, 15623–15628. [Google Scholar] [CrossRef] [Green Version]
  68. Ren, C.; Zhang, J.; Philipson, K.D.; Kotlikoff, M.I.; Blaustein, M.P.; Matteson, D.R. Activation of L-type Ca2+ channels by protein kinase C is reduced in smooth muscle-specific Na+/Ca2+ exchanger knockout mice. Am. J. Physiol. Heart Circ. Physiol. 2010, 298, H1484–H1491. [Google Scholar] [CrossRef] [Green Version]
  69. Weiss, S.; Keren-Raifman, T.; Oz, S.; Ben Mocha, A.; Haase, H.; Dascal, N. Modulation of distinct isoforms of L-type calcium channels by G(q)-coupled receptors in Xenopus oocytes: Antagonistic effects of Gbetagamma and protein kinase C. Channels 2012, 6, 426–437. [Google Scholar] [CrossRef] [Green Version]
  70. Gulia, J.; Navedo, M.F.; Gui, P.; Chao, J.T.; Mercado, J.L.; Santana, L.F.; Davis, M.J. Regulation of L-type calcium channel sparklet activity by c-Src and PKC-alpha. Am. J. Physiol. Cell Physiol. 2013, 305, C568–C577. [Google Scholar] [CrossRef]
  71. Navedo, M.F.; Nieves-Cintron, M.; Amberg, G.C.; Yuan, C.; Votaw, V.S.; Lederer, W.J.; McKnight, G.S.; Santana, L.F. AKAP150 is required for stuttering persistent Ca2+ sparklets and angiotensin II-induced hypertension. Circ. Res. 2008, 102, e1–e11. [Google Scholar] [CrossRef] [Green Version]
  72. Hu, X.Q.; Singh, N.; Mukhopadhyay, D.; Akbarali, H.I. Modulation of voltage-dependent Ca2+ channels in rabbit colonic smooth muscle cells by c-Src and focal adhesion kinase. J. Biol. Chem. 1998, 273, 5337–5342. [Google Scholar] [CrossRef] [Green Version]
  73. Gui, P.; Chao, J.T.; Wu, X.; Yang, Y.; Davis, G.E.; Davis, M.J. Coordinated regulation of vascular Ca2+ and K+ channels by integrin signaling. Adv. Exp. Med. Biol. 2010, 674, 69–79. [Google Scholar] [CrossRef]
  74. Wijetunge, S.; Hughes, A.D. pp60c-src increases voltage-operated calcium channel currents in vascular smooth muscle cells. Biochem. Biophys. Res. Commun. 1995, 217, 1039–1044. [Google Scholar] [CrossRef]
  75. Wijetunge, S.; Hughes, A.D. Activation of endogenous c-Src or a related tyrosine kinase by intracellular (pY)EEI peptide increases voltage-operated calcium channel currents in rabbit ear artery cells. FEBS Lett. 1996, 399, 63–66. [Google Scholar] [CrossRef] [Green Version]
  76. Wijetunge, S.; Lymn, J.S.; Hughes, A.D. Effects of protein tyrosine kinase inhibitors on voltage-operated calcium channel currents in vascular smooth muscle cells and pp60(c-src) kinase activity. Br. J. Pharmacol. 2000, 129, 1347–1354. [Google Scholar] [CrossRef] [Green Version]
  77. Macrez, N.; Mironneau, C.; Carricaburu, V.; Quignard, J.F.; Babich, A.; Czupalla, C.; Nurnberg, B.; Mironneau, J. Phosphoinositide 3-kinase isoforms selectively couple receptors to vascular L-type Ca2+ channels. Circ. Res. 2001, 89, 692–699. [Google Scholar] [CrossRef] [Green Version]
  78. Pinho, J.F.; Medeiros, M.A.; Capettini, L.S.; Rezende, B.A.; Campos, P.P.; Andrade, S.P.; Cortes, S.F.; Cruz, J.S.; Lemos, V.S. Phosphatidylinositol 3-kinase-delta up-regulates L-type Ca2+ currents and increases vascular contractility in a mouse model of type 1 diabetes. Br. J. Pharmacol. 2010, 161, 1458–1471. [Google Scholar] [CrossRef] [Green Version]
  79. Le Blanc, C.; Mironneau, C.; Barbot, C.; Henaff, M.; Bondeva, T.; Wetzker, R.; Macrez, N. Regulation of vascular L-type Ca2+ channels by phosphatidylinositol 3,4,5-trisphosphate. Circ. Res. 2004, 95, 300–307. [Google Scholar] [CrossRef] [Green Version]
  80. Viard, P.; Butcher, A.J.; Halet, G.; Davies, A.; Nurnberg, B.; Heblich, F.; Dolphin, A.C. PI3K promotes voltage-dependent calcium channel trafficking to the plasma membrane. Nat. Neurosci. 2004, 7, 939–946. [Google Scholar] [CrossRef]
  81. Catalucci, D.; Zhang, D.H.; DeSantiago, J.; Aimond, F.; Barbara, G.; Chemin, J.; Bonci, D.; Picht, E.; Rusconi, F.; Dalton, N.D.; et al. Akt regulates L-type Ca2+ channel activity by modulating Cavalpha1 protein stability. J. Cell Biol. 2009, 184, 923–933. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Carnevale, D.; Vecchione, C.; Mascio, G.; Esposito, G.; Cifelli, G.; Martinello, K.; Landolfi, A.; Selvetella, G.; Grieco, P.; Damato, A.; et al. PI3Kgamma inhibition reduces blood pressure by a vasorelaxant Akt/L-type calcium channel mechanism. Cardiovasc. Res. 2012, 93, 200–209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Davis, M.J.; Wu, X.; Nurkiewicz, T.R.; Kawasaki, J.; Davis, G.E.; Hill, M.A.; Meininger, G.A. Integrins and mechanotransduction of the vascular myogenic response. Am. J. Physiol. Heart Circ. Physiol. 2001, 280, H1427–H1433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Wu, X.; Davis, G.E.; Meininger, G.A.; Wilson, E.; Davis, M.J. Regulation of the L-type calcium channel by alpha 5beta 1 integrin requires signaling between focal adhesion proteins. J. Biol. Chem. 2001, 276, 30285–30292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Gui, P.; Wu, X.; Ling, S.; Stotz, S.C.; Winkfein, R.J.; Wilson, E.; Davis, G.E.; Braun, A.P.; Zamponi, G.W.; Davis, M.J. Integrin receptor activation triggers converging regulation of Cav1.2 calcium channels by c-Src and protein kinase A pathways. J. Biol. Chem. 2006, 281, 14015–14025. [Google Scholar] [CrossRef] [Green Version]
  86. Waitkus-Edwards, K.R.; Martinez-Lemus, L.A.; Wu, X.; Trzeciakowski, J.P.; Davis, M.J.; Davis, G.E.; Meininger, G.A. alpha(4)beta(1) Integrin activation of L-type calcium channels in vascular smooth muscle causes arteriole vasoconstriction. Circ. Res. 2002, 90, 473–480. [Google Scholar] [CrossRef] [Green Version]
  87. Correll, R.N.; Pang, C.; Niedowicz, D.M.; Finlin, B.S.; Andres, D.A. The RGK family of GTP-binding proteins: Regulators of voltage-dependent calcium channels and cytoskeleton remodeling. Cell. Signal. 2008, 20, 292–300. [Google Scholar] [CrossRef] [Green Version]
  88. Yang, T.; Colecraft, H.M. Regulation of voltage-dependent calcium channels by RGK proteins. Biochim. Biophys. Acta 2013, 1828, 1644–1654. [Google Scholar] [CrossRef] [Green Version]
  89. Katchman, A.; Yang, L.; Zakharov, S.I.; Kushner, J.; Abrams, J.; Chen, B.X.; Liu, G.; Pitt, G.S.; Colecraft, H.M.; Marx, S.O. Proteolytic cleavage and PKA phosphorylation of alpha1C subunit are not required for adrenergic regulation of Cav1.2 in the heart. Proc. Natl. Acad. Sci. USA 2017, 114, 9194–9199. [Google Scholar] [CrossRef] [Green Version]
  90. Liu, G.; Papa, A.; Katchman, A.N.; Zakharov, S.I.; Roybal, D.; Hennessey, J.A.; Kushner, J.; Yang, L.; Chen, B.X.; Kushnir, A.; et al. Mechanism of adrenergic Cav1.2 stimulation revealed by proximity proteomics. Nature 2020, 577, 695–700. [Google Scholar] [CrossRef]
  91. Yang, T.; Puckerin, A.; Colecraft, H.M. Distinct RGK GTPases differentially use alpha1- and auxiliary beta-binding-dependent mechanisms to inhibit Cav1.2/Cav2.2 channels. PLoS ONE 2012, 7, e37079. [Google Scholar] [CrossRef]
  92. Puckerin, A.A.; Chang, D.D.; Shuja, Z.; Choudhury, P.; Scholz, J.; Colecraft, H.M. Engineering selectivity into RGK GTPase inhibition of voltage-dependent calcium channels. Proc. Natl. Acad. Sci. USA 2018, 115, 12051–12056. [Google Scholar] [CrossRef] [Green Version]
  93. Beguin, P.; Nagashima, K.; Gonoi, T.; Shibasaki, T.; Takahashi, K.; Kashima, Y.; Ozaki, N.; Geering, K.; Iwanaga, T.; Seino, S. Regulation of Ca2+ channel expression at the cell surface by the small G-protein kir/Gem. Nature 2001, 411, 701–706. [Google Scholar] [CrossRef]
  94. Finlin, B.S.; Crump, S.M.; Satin, J.; Andres, D.A. Regulation of voltage-gated calcium channel activity by the Rem and Rad GTPases. Proc. Natl. Acad. Sci. USA 2003, 100, 14469–14474. [Google Scholar] [CrossRef] [Green Version]
  95. Xu, X.; Marx, S.O.; Colecraft, H.M. Molecular mechanisms, and selective pharmacological rescue, of Rem-inhibited Cav1.2 channels in heart. Circ. Res. 2010, 107, 620–630. [Google Scholar] [CrossRef] [Green Version]
  96. Bannister, J.P.; Bulley, S.; Leo, M.D.; Kidd, M.W.; Jaggar, J.H. Rab25 influences functional Cav1.2 channel surface expression in arterial smooth muscle cells. Am. J. Physiol. Cell Physiol. 2016, 310, C885–C893. [Google Scholar] [CrossRef] [Green Version]
  97. Davis, M.J.; Hill, M.A. Signaling mechanisms underlying the vascular myogenic response. Physiol. Rev. 1999, 79, 387–423. [Google Scholar] [CrossRef] [Green Version]
  98. Harder, D.R. Pressure-dependent membrane depolarization in cat middle cerebral artery. Circ. Res. 1984, 55, 197–202. [Google Scholar] [CrossRef] [Green Version]
  99. Knot, H.J.; Nelson, M.T. Regulation of arterial diameter and wall [Ca2+] in cerebral arteries of rat by membrane potential and intravascular pressure. J. Physiol. 1998, 508 Pt 1, 199–209. [Google Scholar] [CrossRef]
  100. Kotecha, N.; Hill, M.A. Myogenic contraction in rat skeletal muscle arterioles: Smooth muscle membrane potential and Ca2+ signaling. Am. J. Physiol. Heart Circ. Physiol. 2005, 289, H1326–H1334. [Google Scholar] [CrossRef]
  101. Welsh, D.G.; Morielli, A.D.; Nelson, M.T.; Brayden, J.E. Transient receptor potential channels regulate myogenic tone of resistance arteries. Circ. Res. 2002, 90, 248–250. [Google Scholar] [CrossRef] [PubMed]
  102. Earley, S.; Waldron, B.J.; Brayden, J.E. Critical role for transient receptor potential channel TRPM4 in myogenic constriction of cerebral arteries. Circ. Res. 2004, 95, 922–929. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Sharif-Naeini, R.; Dedman, A.; Folgering, J.H.; Duprat, F.; Patel, A.; Nilius, B.; Honore, E. TRP channels and mechanosensory transduction: Insights into the arterial myogenic response. Pflügers Arch. 2008, 456, 529–540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Kim, E.C.; Choi, S.K.; Lim, M.; Yeon, S.I.; Lee, Y.H. Role of endogenous ENaC and TRP channels in the myogenic response of rat posterior cerebral arteries. PLoS ONE 2013, 8, e84194. [Google Scholar] [CrossRef]
  105. Nemeth, Z.; Hildebrandt, E.; Ryan, M.J.; Granger, J.P.; Drummond, H.A. Pressure-induced constriction of the middle cerebral artery is abolished in TrpC6 knockout mice. Am. J. Physiol. Heart Circ. Physiol. 2020, 319, H42–H50. [Google Scholar] [CrossRef]
  106. Jackson, W.F. Myogenic Tone in Peripheral Resistance Arteries and Arterioles: The Pressure Is On! Front. Physiol. 2021, 12, 699517. [Google Scholar] [CrossRef]
  107. Skaik, K.; Shahzad, M. The emerging role of TRPV1 in myogenic tone. J. Physiol. 2022, 600, 2287–2288. [Google Scholar] [CrossRef]
  108. Nilsson, H.; Jensen, P.E.; Mulvany, M.J. Minor role for direct adrenoceptor-mediated calcium entry in rat mesenteric small arteries. J. Vasc. Res. 1994, 31, 314–321. [Google Scholar] [CrossRef]
  109. Wesselman, J.P.; VanBavel, E.; Pfaffendorf, M.; Spaan, J.A. Voltage-operated calcium channels are essential for the myogenic responsiveness of cannulated rat mesenteric small arteries. J. Vasc. Res. 1996, 33, 32–41. [Google Scholar] [CrossRef]
  110. Miller, F.J., Jr.; Dellsperger, K.C.; Gutterman, D.D. Myogenic constriction of human coronary arterioles. Am. J. Physiol. 1997, 273, H257–H264. [Google Scholar] [CrossRef]
  111. Potocnik, S.J.; Murphy, T.V.; Kotecha, N.; Hill, M.A. Effects of mibefradil and nifedipine on arteriolar myogenic responsiveness and intracellular Ca2+. Br. J. Pharmacol. 2000, 131, 1065–1072. [Google Scholar] [CrossRef] [Green Version]
  112. Murphy, T.V.; Spurrell, B.E.; Hill, M.A. Mechanisms underlying pervanadate-induced contraction of rat cremaster muscle arterioles. Eur. J. Pharmacol. 2002, 442, 107–114. [Google Scholar] [CrossRef]
  113. Ahmed, A.; Waters, C.M.; Leffler, C.W.; Jaggar, J.H. Ionic mechanisms mediating the myogenic response in newborn porcine cerebral arteries. Am. J. Physiol. Heart Circ. Physiol. 2004, 287, H2061–H2069. [Google Scholar] [CrossRef] [Green Version]
  114. Abd El-Rahman, R.R.; Harraz, O.F.; Brett, S.E.; Anfinogenova, Y.; Mufti, R.E.; Goldman, D.; Welsh, D.G. Identification of L- and T-type Ca2+ channels in rat cerebral arteries: Role in myogenic tone development. Am. J. Physiol. Heart Circ. Physiol. 2013, 304, H58–H71. [Google Scholar] [CrossRef] [Green Version]
  115. Jackson, W.F.; Boerman, E.M. Voltage-gated Ca2+ channel activity modulates smooth muscle cell calcium waves in hamster cremaster arterioles. Am. J. Physiol. Heart Circ. Physiol. 2018, 315, H871–H878. [Google Scholar] [CrossRef] [Green Version]
  116. Moosmang, S.; Schulla, V.; Welling, A.; Feil, R.; Feil, S.; Wegener, J.W.; Hofmann, F.; Klugbauer, N. Dominant role of smooth muscle L-type calcium channel Cav1.2 for blood pressure regulation. EMBO J. 2003, 22, 6027–6034. [Google Scholar] [CrossRef]
  117. Fernandez, J.A.; McGahon, M.K.; McGeown, J.G.; Curtis, T.M. Cav3.1 T-Type Ca2+ Channels Contribute to Myogenic Signaling in Rat Retinal Arterioles. Investig. Ophthalmol. Vis. Sci. 2015, 56, 5125–5132. [Google Scholar] [CrossRef] [Green Version]
  118. Chao, J.T.; Gui, P.; Zamponi, G.W.; Davis, G.E.; Davis, M.J. Spatial association of the Cav1.2 calcium channel with alpha5beta1-integrin. Am. J. Physiol. Cell Physiol. 2011, 300, C477–C489. [Google Scholar] [CrossRef] [Green Version]
  119. Mederos y Schnitzler, M.; Storch, U.; Meibers, S.; Nurwakagari, P.; Breit, A.; Essin, K.; Gollasch, M.; Gudermann, T. Gq-coupled receptors as mechanosensors mediating myogenic vasoconstriction. EMBO J. 2008, 27, 3092–3103. [Google Scholar] [CrossRef]
  120. Jackson, W.F.; Boerman, E.M. Regional heterogeneity in the mechanisms of myogenic tone in hamster arterioles. Am. J. Physiol. Heart Circ. Physiol. 2017, 313, H667–H675. [Google Scholar] [CrossRef]
  121. Hill, M.A.; Yang, Y.; Ella, S.R.; Davis, M.J.; Braun, A.P. Large conductance, Ca2+-activated K+ channels (BKCa) and arteriolar myogenic signaling. FEBS Lett. 2010, 584, 2033–2042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Jackson, W.F. Ion channels and the regulation of myogenic tone in peripheral arterioles. Curr. Top. Membr. 2020, 85, 19–58. [Google Scholar] [CrossRef] [PubMed]
  123. Jackson, W.F. Calcium-Dependent Ion Channels and the Regulation of Arteriolar Myogenic Tone. Front. Physiol. 2021, 12, 770450. [Google Scholar] [CrossRef]
  124. Wallis, S.J.; Firth, J.; Dunn, W.R. Pressure-induced myogenic responses in human isolated cerebral resistance arteries. Stroke 1996, 27, 2287–2290, discussion 2291. [Google Scholar] [CrossRef]
  125. Falcone, J.C.; Davis, M.J.; Meininger, G.A. Endothelial independence of myogenic response in isolated skeletal muscle arterioles. Am. J. Physiol. 1991, 260, H130–H135. [Google Scholar] [CrossRef]
  126. Kuo, L.; Chilian, W.M.; Davis, M.J. Coronary arteriolar myogenic response is independent of endothelium. Circ. Res. 1990, 66, 860–866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Bagher, P.; Beleznai, T.; Kansui, Y.; Mitchell, R.; Garland, C.J.; Dora, K.A. Low intravascular pressure activates endothelial cell TRPV4 channels, local Ca2+ events, and IKCa channels, reducing arteriolar tone. Proc. Natl. Acad. Sci. USA 2012, 109, 18174–18179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Pires, P.W.; Sullivan, M.N.; Pritchard, H.A.; Robinson, J.J.; Earley, S. Unitary TRPV3 channel Ca2+ influx events elicit endothelium-dependent dilation of cerebral parenchymal arterioles. Am. J. Physiol. Heart Circ. Physiol. 2015, 309, H2031–H2041. [Google Scholar] [CrossRef] [Green Version]
  129. Veerareddy, S.; Cooke, C.L.; Baker, P.N.; Davidge, S.T. Vascular adaptations to pregnancy in mice: Effects on myogenic tone. Am. J. Physiol. Heart Circ. Physiol. 2002, 283, H2226–H2233. [Google Scholar] [CrossRef] [Green Version]
  130. Carnevale, D.; Facchinello, N.; Iodice, D.; Bizzotto, D.; Perrotta, M.; De Stefani, D.; Pallante, F.; Carnevale, L.; Ricciardi, F.; Cifelli, G.; et al. Loss of EMILIN-1 Enhances Arteriolar Myogenic Tone Through TGF-beta (Transforming Growth Factor-beta)-Dependent Transactivation of EGFR (Epidermal Growth Factor Receptor) and Is Relevant for Hypertension in Mice and Humans. Arterioscler. Thromb. Vasc. Biol. 2018, 38, 2484–2497. [Google Scholar] [CrossRef]
  131. Ren, Y.; D’Ambrosio, M.A.; Liu, R.; Pagano, P.J.; Garvin, J.L.; Carretero, O.A. Enhanced myogenic response in the afferent arteriole of spontaneously hypertensive rats. Am. J. Physiol. Heart Circ. Physiol. 2010, 298, H1769–H1775. [Google Scholar] [CrossRef]
  132. Nademi, S.; Lu, C.; Dickhout, J.G. Enhanced Myogenic Constriction in the SHR Preglomerular Vessels Is Mediated by Thromboxane A2 Synthesis. Front. Physiol. 2020, 11, 853. [Google Scholar] [CrossRef]
  133. Izzard, A.S.; Bund, S.J.; Heagerty, A.M. Myogenic tone in mesenteric arteries from spontaneously hypertensive rats. Am. J. Physiol. 1996, 270, H1–H6. [Google Scholar] [CrossRef]
  134. Linde, C.I.; Karashima, E.; Raina, H.; Zulian, A.; Wier, W.G.; Hamlyn, J.M.; Ferrari, P.; Blaustein, M.P.; Golovina, V.A. Increased arterial smooth muscle Ca2+ signaling, vasoconstriction, and myogenic reactivity in Milan hypertensive rats. Am. J. Physiol. Heart Circ. Physiol. 2012, 302, H611–H620. [Google Scholar] [CrossRef] [Green Version]
  135. Dunn, W.R.; Wallis, S.J.; Gardiner, S.M. Remodelling and enhanced myogenic tone in cerebral resistance arteries isolated from genetically hypertensive Brattleboro rats. J. Vasc. Res. 1998, 35, 18–26. [Google Scholar] [CrossRef]
  136. Jarajapu, Y.P.; Knot, H.J. Relative contribution of Rho kinase and protein kinase C to myogenic tone in rat cerebral arteries in hypertension. Am. J. Physiol. Heart Circ. Physiol. 2005, 289, H1917–H1922. [Google Scholar] [CrossRef] [Green Version]
  137. Ahn, D.S.; Choi, S.K.; Kim, Y.H.; Cho, Y.E.; Shin, H.M.; Morgan, K.G.; Lee, Y.H. Enhanced stretch-induced myogenic tone in the basilar artery of spontaneously hypertensive rats. J. Vasc. Res. 2007, 44, 182–191. [Google Scholar] [CrossRef]
  138. Gonzalez, J.M.; Somoza, B.; Conde, M.V.; Fernandez-Alfonso, M.S.; Gonzalez, M.C.; Arribas, S.M. Hypertension increases middle cerebral artery resting tone in spontaneously hypertensive rats: Role of tonic vasoactive factor availability. Clin. Sci. 2008, 114, 651–659. [Google Scholar] [CrossRef] [Green Version]
  139. Falcone, J.C.; Granger, H.J.; Meininger, G.A. Enhanced myogenic activation in skeletal muscle arterioles from spontaneously hypertensive rats. Am. J. Physiol. 1993, 265, H1847–H1855. [Google Scholar] [CrossRef]
  140. Shibuya, J.; Ohyanagi, M.; Iwasaki, T. Enhanced myogenic response in resistance small arteries from spontaneously hypertensive rats: Relationship to the voltage-dependent calcium channel. Am. J. Hypertens. 1998, 11, 767–773. [Google Scholar] [CrossRef]
  141. Toth, P.; Csiszar, A.; Tucsek, Z.; Sosnowska, D.; Gautam, T.; Koller, A.; Schwartzman, M.L.; Sonntag, W.E.; Ungvari, Z. Role of 20-HETE, TRPC channels, and BKCa in dysregulation of pressure-induced Ca2+ signaling and myogenic constriction of cerebral arteries in aged hypertensive mice. Am. J. Physiol. Heart Circ. Physiol. 2013, 305, H1698–H1708. [Google Scholar] [CrossRef] [Green Version]
  142. McCurley, A.; Pires, P.W.; Bender, S.B.; Aronovitz, M.; Zhao, M.J.; Metzger, D.; Chambon, P.; Hill, M.A.; Dorrance, A.M.; Mendelsohn, M.E.; et al. Direct regulation of blood pressure by smooth muscle cell mineralocorticoid receptors. Nat. Med. 2012, 18, 1429–1433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. DuPont, J.J.; McCurley, A.; Davel, A.P.; McCarthy, J.; Bender, S.B.; Hong, K.; Yang, Y.; Yoo, J.K.; Aronovitz, M.; Baur, W.E.; et al. Vascular mineralocorticoid receptor regulates microRNA-155 to promote vasoconstriction and rising blood pressure with aging. JCI Insight 2016, 1, e88942. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Sauve, M.; Hui, S.K.; Dinh, D.D.; Foltz, W.D.; Momen, A.; Nedospasov, S.A.; Offermanns, S.; Husain, M.; Kroetsch, J.T.; Lidington, D.; et al. Tumor Necrosis Factor/Sphingosine-1-Phosphate Signaling Augments Resistance Artery Myogenic Tone in Diabetes. Diabetes 2016, 65, 1916–1928. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. King, A.; Bowe, J. Animal models for diabetes: Understanding the pathogenesis and finding new treatments. Biochem. Pharmacol. 2016, 99, 1–10. [Google Scholar] [CrossRef]
  146. Kottaisamy, C.P.D.; Raj, D.S.; Prasanth Kumar, V.; Sankaran, U. Experimental animal models for diabetes and its related complications—A review. Lab. Anim. Res. 2021, 37, 23. [Google Scholar] [CrossRef]
  147. Bunag, R.D.; Tomita, T.; Sasaki, S. Streptozotocin diabetic rats are hypertensive despite reduced hypothalamic responsiveness. Hypertension 1982, 4, 556–565. [Google Scholar] [CrossRef] [Green Version]
  148. Bagi, Z.; Erdei, N.; Toth, A.; Li, W.; Hintze, T.H.; Koller, A.; Kaley, G. Type 2 diabetic mice have increased arteriolar tone and blood pressure: Enhanced release of COX-2-derived constrictor prostaglandins. Arterioscler. Thromb. Vasc. Biol. 2005, 25, 1610–1616. [Google Scholar] [CrossRef] [Green Version]
  149. Senador, D.; Kanakamedala, K.; Irigoyen, M.C.; Morris, M.; Elased, K.M. Cardiovascular and autonomic phenotype of db/db diabetic mice. Exp. Physiol. 2009, 94, 648–658. [Google Scholar] [CrossRef] [Green Version]
  150. Bhandari, U.; Kumar, V.; Khanna, N.; Panda, B.P. The effect of high-fat diet-induced obesity on cardiovascular toxicity in Wistar albino rats. Hum. Exp. Toxicol. 2011, 30, 1313–1321. [Google Scholar] [CrossRef]
  151. Musial, D.C.; da Silva Junior, E.D.; da Silva, R.M.; Miranda-Ferreira, R.; Lima-Landman, M.T.; Jurkiewicz, A.; Garcia, A.G.; Jurkiewicz, N.H. Increase of angiotensin-converting enzyme activity and peripheral sympathetic dysfunction could contribute to hypertension development in streptozotocin-induced diabetic rats. Diab. Vasc. Dis. Res. 2013, 10, 498–504. [Google Scholar] [CrossRef] [Green Version]
  152. Alameddine, A.; Fajloun, Z.; Bourreau, J.; Gauquelin-Koch, G.; Yuan, M.; Gauguier, D.; Derbre, S.; Ayer, A.; Custaud, M.A.; Navasiolava, N. The cardiovascular effects of salidroside in the Goto-Kakizaki diabetic rat model. J. Physiol. Pharmacol. 2015, 66, 249–257. [Google Scholar]
  153. Ma, Y.G.; Wang, J.W.; Bai, Y.G.; Liu, M.; Xie, M.J.; Dai, Z.J. Salidroside contributes to reducing blood pressure and alleviating cerebrovascular contractile activity in diabetic Goto-Kakizaki Rats by inhibition of L-type calcium channel in smooth muscle cells. BMC Pharmacol. Toxicol. 2017, 18, 30. [Google Scholar] [CrossRef] [Green Version]
  154. Wang, Y.W.; Yu, H.R.; Tiao, M.M.; Tain, Y.L.; Lin, I.C.; Sheen, J.M.; Lin, Y.J.; Chang, K.A.; Chen, C.C.; Tsai, C.C.; et al. Maternal Obesity Related to High Fat Diet Induces Placenta Remodeling and Gut Microbiome Shaping That Are Responsible for Fetal Liver Lipid Dysmetabolism. Front. Nutr. 2021, 8, 736944. [Google Scholar] [CrossRef]
  155. Carrillo-Sepulveda, M.A.; Maddie, N.; Johnson, C.M.; Burke, C.; Lutz, O.; Yakoub, B.; Kramer, B.; Persand, D. Vascular hyperacetylation is associated with vascular smooth muscle dysfunction in a rat model of non-obese type 2 diabetes. Mol. Med. 2022, 28, 30. [Google Scholar] [CrossRef]
  156. Choi, S.K.; Kwon, Y.; Byeon, S.; Lee, Y.H. Stimulation of autophagy improves vascular function in the mesenteric arteries of type 2 diabetic mice. Exp. Physiol. 2020, 105, 192–200. [Google Scholar] [CrossRef]
  157. Zimmermann, P.A.; Knot, H.J.; Stevenson, A.S.; Nelson, M.T. Increased myogenic tone and diminished responsiveness to ATP-sensitive K+ channel openers in cerebral arteries from diabetic rats. Circ. Res. 1997, 81, 996–1004. [Google Scholar] [CrossRef]
  158. Jarajapu, Y.P.; Guberski, D.L.; Grant, M.B.; Knot, H.J. Myogenic tone and reactivity of cerebral arteries in type II diabetic BBZDR/Wor rat. Eur. J. Pharmacol. 2008, 579, 298–307. [Google Scholar] [CrossRef] [Green Version]
  159. Ungvari, Z.; Pacher, P.; Kecskemeti, V.; Papp, G.; Szollar, L.; Koller, A. Increased myogenic tone in skeletal muscle arterioles of diabetic rats. Possible role of increased activity of smooth muscle Ca2+ channels and protein kinase C. Cardiovasc. Res. 1999, 43, 1018–1028. [Google Scholar] [CrossRef] [Green Version]
  160. Ito, I.; Jarajapu, Y.P.; Guberski, D.L.; Grant, M.B.; Knot, H.J. Myogenic tone and reactivity of rat ophthalmic artery in acute exposure to high glucose and in a type II diabetic model. Investig. Ophthalmol. Vis. Sci. 2006, 47, 683–692. [Google Scholar] [CrossRef] [Green Version]
  161. Prada, M.P.; Syed, A.U.; Buonarati, O.R.; Reddy, G.R.; Nystoriak, M.A.; Ghosh, D.; Simo, S.; Sato, D.; Sasse, K.C.; Ward, S.M.; et al. A Gs-coupled purinergic receptor boosts Ca2+ influx and vascular contractility during diabetic hyperglycemia. Elife 2019, 8, e42214. [Google Scholar] [CrossRef] [PubMed]
  162. Kublickiene, K.R.; Lindblom, B.; Kruger, K.; Nisell, H. Preeclampsia: Evidence for impaired shear stress-mediated nitric oxide release in uterine circulation. Am. J. Obstet. Gynecol. 2000, 183, 160–166. [Google Scholar] [CrossRef] [PubMed]
  163. Alexander, B.T.; Kassab, S.E.; Miller, M.T.; Abram, S.R.; Reckelhoff, J.F.; Bennett, W.A.; Granger, J.P. Reduced uterine perfusion pressure during pregnancy in the rat is associated with increases in arterial pressure and changes in renal nitric oxide. Hypertension 2001, 37, 1191–1195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Fushima, T.; Sekimoto, A.; Minato, T.; Ito, T.; Oe, Y.; Kisu, K.; Sato, E.; Funamoto, K.; Hayase, T.; Kimura, Y.; et al. Reduced Uterine Perfusion Pressure (RUPP) Model of Preeclampsia in Mice. PLoS ONE 2016, 11, e0155426. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Ramirez, R.J.; Debrah, J.; Novak, J. Increased myogenic responses of resistance-sized mesenteric arteries after reduced uterine perfusion pressure in pregnant rats. Hypertens. Pregnancy 2011, 30, 45–57. [Google Scholar] [CrossRef]
  166. Reho, J.J.; Peck, J.; Novak, J.; Ramirez, R.J. Hypertension induced by episodic reductions in uteroplacental blood flow in gravid rat. Hypertens. Pregnancy 2011, 30, 208–220. [Google Scholar] [CrossRef]
  167. Reho, J.J.; Toot, J.D.; Peck, J.; Novak, J.; Yun, Y.H.; Ramirez, R.J. Increased Myogenic Reactivity of Uterine Arteries from Pregnant Rats with Reduced Uterine Perfusion Pressure. Pregnancy Hypertens. 2012, 2, 106–114. [Google Scholar] [CrossRef] [Green Version]
  168. Powell, J.S.; Gandley, R.E.; Lackner, E.; Dolinish, A.; Ouyang, Y.; Powers, R.W.; Morelli, A.E.; Hubel, C.A.; Sadovsky, Y. Small extracellular vesicles from plasma of women with preeclampsia increase myogenic tone and decrease endothelium-dependent relaxation of mouse mesenteric arteries. Pregnancy Hypertens. 2022, 28, 66–73. [Google Scholar] [CrossRef]
  169. Soleymanlou, N.; Jurisica, I.; Nevo, O.; Ietta, F.; Zhang, X.; Zamudio, S.; Post, M.; Caniggia, I. Molecular evidence of placental hypoxia in preeclampsia. J. Clin. Endocrinol. Metab. 2005, 90, 4299–4308. [Google Scholar] [CrossRef] [Green Version]
  170. Ducsay, C.A.; Goyal, R.; Pearce, W.J.; Wilson, S.; Hu, X.Q.; Zhang, L. Gestational Hypoxia and Developmental Plasticity. Physiol. Rev. 2018, 98, 1241–1334. [Google Scholar] [CrossRef]
  171. Tong, W.; Giussani, D.A. Preeclampsia link to gestational hypoxia. J. Dev. Orig. Health Dis. 2019, 10, 322–333. [Google Scholar] [CrossRef]
  172. Grant, I.D.; Giussani, D.A.; Aiken, C.E. Blood pressure and hypertensive disorders of pregnancy at high altitude: A systematic review and meta-analysis. Am. J. Obstet. Gynecol. MFM 2021, 3, 100400. [Google Scholar] [CrossRef]
  173. Palmer, S.K.; Moore, L.G.; Young, D.; Cregger, B.; Berman, J.C.; Zamudio, S. Altered blood pressure course during normal pregnancy and increased preeclampsia at high altitude (3100 m) in Colorado. Am. J. Obstet. Gynecol. 1999, 180, 1161–1168. [Google Scholar] [CrossRef]
  174. Keyes, L.E.; Armaza, J.F.; Niermeyer, S.; Vargas, E.; Young, D.A.; Moore, L.G. Intrauterine growth restriction, preeclampsia, and intrauterine mortality at high altitude in Bolivia. Pediatr. Res. 2003, 54, 20–25. [Google Scholar] [CrossRef]
  175. Zamudio, S. High-altitude hypoxia and preeclampsia. Front. Biosci. 2007, 12, 2967–2977. [Google Scholar] [CrossRef]
  176. Hu, X.Q.; Dasgupta, C.; Xiao, D.; Huang, X.; Yang, S.; Zhang, L. MicroRNA-210 Targets Ten-Eleven Translocation Methylcytosine Dioxygenase 1 and Suppresses Pregnancy-Mediated Adaptation of Large Conductance Ca2+-Activated K+ Channel Expression and Function in Ovine Uterine Arteries. Hypertension 2017, 70, 601–612. [Google Scholar] [CrossRef]
  177. Hu, X.Q.; Song, R.; Romero, M.; Dasgupta, C.; Min, J.; Hatcher, D.; Xiao, D.; Blood, A.; Wilson, S.M.; Zhang, L. Gestational Hypoxia Inhibits Pregnancy-Induced Upregulation of Ca2+ Sparks and Spontaneous Transient Outward Currents in Uterine Arteries Via Heightened Endoplasmic Reticulum/Oxidative Stress. Hypertension 2020, 76, 930–942. [Google Scholar] [CrossRef]
  178. Tong, W.; Allison, B.J.; Brain, K.L.; Patey, O.V.; Niu, Y.; Botting, K.J.; Ford, S.G.; Garrud, T.A.; Wooding, P.F.B.; Shaw, C.J.; et al. Chronic Hypoxia in Ovine Pregnancy Recapitulates Physiological and Molecular Markers of Preeclampsia in the Mother, Placenta, and Offspring. Hypertension 2022, 79, 1525–1535. [Google Scholar] [CrossRef]
  179. Chang, K.; Xiao, D.; Huang, X.; Longo, L.D.; Zhang, L. Chronic hypoxia increases pressure-dependent myogenic tone of the uterine artery in pregnant sheep: Role of ERK/PKC pathway. Am. J. Physiol. Heart Circ. Physiol. 2009, 296, H1840–H1849. [Google Scholar] [CrossRef] [Green Version]
  180. Hu, X.Q.; Chen, M.; Dasgupta, C.; Xiao, D.; Huang, X.; Yang, S.; Zhang, L. Chronic hypoxia upregulates DNA methyltransferase and represses large conductance Ca2+-activated K+ channel function in ovine uterine arteries. Biol. Reprod. 2017, 96, 424–434. [Google Scholar] [CrossRef] [Green Version]
  181. Pires, P.W.; Jackson, W.F.; Dorrance, A.M. Regulation of myogenic tone and structure of parenchymal arterioles by hypertension and the mineralocorticoid receptor. Am. J. Physiol. Heart Circ. Physiol. 2015, 309, H127–H136. [Google Scholar] [CrossRef] [PubMed]
  182. Tarjus, A.; Belozertseva, E.; Louis, H.; El Moghrabi, S.; Labat, C.; Lacolley, P.; Jaisser, F.; Galmiche, G. Role of smooth muscle cell mineralocorticoid receptor in vascular tone. Pflügers Arch. 2015, 467, 1643–1650. [Google Scholar] [CrossRef] [PubMed]
  183. Callera, G.E.; Varanda, W.A.; Bendhack, L.M. Ca2+ influx is increased in 2-kidney, 1-clip hypertensive rat aorta. Hypertension 2001, 38, 592–596. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Ebeigbe, A.B.; Ezimokhai, M.; Aloamaka, C.P. Responses of arterial smooth muscle from normotensive and pre-eclamptic subjects to the calcium channel agonist, Bay K 8644. Res. Exp. Med. 1987, 187, 461–468. [Google Scholar] [CrossRef] [PubMed]
  185. Matsuda, K.; Lozinskaya, I.; Cox, R.H. Augmented contributions of voltage-gated Ca2+ channels to contractile responses in spontaneously hypertensive rat mesenteric arteries. Am. J. Hypertens. 1997, 10, 1231–1239. [Google Scholar] [CrossRef] [Green Version]
  186. Murphy, J.G.; Fleming, J.B.; Cockrell, K.L.; Granger, J.P.; Khalil, R.A. [Ca2+]i signaling in renal arterial smooth muscle cells of pregnant rat is enhanced during inhibition of NOS. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2001, 280, R87–R99. [Google Scholar] [CrossRef] [Green Version]
  187. Pratt, P.F.; Bonnet, S.; Ludwig, L.M.; Bonnet, P.; Rusch, N.J. Upregulation of L-type Ca2+ channels in mesenteric and skeletal arteries of SHR. Hypertension 2002, 40, 214–219. [Google Scholar] [CrossRef] [Green Version]
  188. Murphy, J.G.; Herrington, J.N.; Granger, J.P.; Khalil, R.A. Enhanced [Ca2+]i in renal arterial smooth muscle cells of pregnant rats with reduced uterine perfusion pressure. Am. J. Physiol. Heart Circ. Physiol. 2003, 284, H393–H403. [Google Scholar] [CrossRef] [Green Version]
  189. White, R.E.; Carrier, G.O. Vascular contraction induced by activation of membrane calcium ion channels is enhanced in streptozotocin-diabetes. J. Pharmacol. Exp. Ther. 1990, 253, 1057–1062. [Google Scholar]
  190. Chen, W.; Khalil, R.A. Differential [Ca2+]i signaling of vasoconstriction in mesenteric microvessels of normal and reduced uterine perfusion pregnant rats. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2008, 295, R1962–R1972. [Google Scholar] [CrossRef] [Green Version]
  191. Aoki, K.; Asano, M. Effects of Bay K 8644 and nifedipine on femoral arteries of spontaneously hypertensive rats. Br. J. Pharmacol. 1986, 88, 221–230. [Google Scholar] [CrossRef] [Green Version]
  192. Bannister, J.P.; Bulley, S.; Narayanan, D.; Thomas-Gatewood, C.; Luzny, P.; Pachuau, J.; Jaggar, J.H. Transcriptional upregulation of alpha2delta-1 elevates arterial smooth muscle cell voltage-dependent Ca2+ channel surface expression and cerebrovascular constriction in genetic hypertension. Hypertension 2012, 60, 1006–1015. [Google Scholar] [CrossRef]
  193. Liao, J.; Zhang, Y.; Ye, F.; Zhang, L.; Chen, Y.; Zeng, F.; Shi, L. Epigenetic regulation of L-type voltage-gated Ca2+ channels in mesenteric arteries of aging hypertensive rats. Hypertens. Res. 2017, 40, 441–449. [Google Scholar] [CrossRef]
  194. Cox, R.H.; Lozinskaya, I.M. Augmented calcium currents in mesenteric artery branches of the spontaneously hypertensive rat. Hypertension 1995, 26, 1060–1064. [Google Scholar] [CrossRef]
  195. Ohya, Y.; Abe, I.; Fujii, K.; Takata, Y.; Fujishima, M. Voltage-dependent Ca2+ channels in resistance arteries from spontaneously hypertensive rats. Circ. Res. 1993, 73, 1090–1099. [Google Scholar] [CrossRef] [Green Version]
  196. Lozinskaya, I.M.; Cox, R.H. Effects of age on Ca2+ currents in small mesenteric artery myocytes from Wistar-Kyoto and spontaneously hypertensive rats. Hypertension 1997, 29, 1329–1336. [Google Scholar] [CrossRef]
  197. Simard, J.M.; Li, X.; Tewari, K. Increase in functional Ca2+ channels in cerebral smooth muscle with renal hypertension. Circ. Res. 1998, 82, 1330–1337. [Google Scholar] [CrossRef] [Green Version]
  198. Takimoto, K.; Li, D.; Nerbonne, J.M.; Levitan, E.S. Distribution, splicing and glucocorticoid-induced expression of cardiac alpha 1C and alpha 1D voltage-gated Ca2+ channel mRNAs. J. Mol. Cell. Cardiol. 1997, 29, 3035–3042. [Google Scholar] [CrossRef]
  199. Obejero-Paz, C.A.; Lakshmanan, M.; Jones, S.W.; Scarpa, A. Effects of dexamethasone on L-type calcium currents in the A7r5 smooth muscle-derived cell line. FEBS Lett. 1993, 333, 73–77. [Google Scholar] [CrossRef] [Green Version]
  200. Yu, Z.; Wang, T.; Xu, L.; Huang, C.X. Thyroid hormone increased L-type calcium channel mRNA expression and L-type calcium current of myocytes in rabbits. Biomed. Mater. Eng. 2012, 22, 49–55. [Google Scholar] [CrossRef]
  201. Wang, J.; Thio, S.S.; Yang, S.S.; Yu, D.; Yu, C.Y.; Wong, Y.P.; Liao, P.; Li, S.; Soong, T.W. Splice variant specific modulation of Cav1.2 calcium channel by galectin-1 regulates arterial constriction. Circ. Res. 2011, 109, 1250–1258. [Google Scholar] [CrossRef] [Green Version]
  202. Hu, Z.; Li, G.; Wang, J.W.; Chong, S.Y.; Yu, D.; Wang, X.; Soon, J.L.; Liang, M.C.; Wong, Y.P.; Huang, N.; et al. Regulation of Blood Pressure by Targeting Cav1.2-Galectin-1 Protein Interaction. Circulation 2018, 138, 1431–1445. [Google Scholar] [CrossRef] [PubMed]
  203. Ma, Y.G.; Zhang, Y.B.; Bai, Y.G.; Dai, Z.J.; Liang, L.; Liu, M.; Xie, M.J.; Guan, H.T. Berberine alleviates the cerebrovascular contractility in streptozotocin-induced diabetic rats through modulation of intracellular Ca2+ handling in smooth muscle cells. Cardiovasc. Diabetol. 2016, 15, 63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Wilde, D.W.; Massey, K.D.; Walker, G.K.; Vollmer, A.; Grekin, R.J. High-fat diet elevates blood pressure and cerebrovascular muscle Ca2+ current. Hypertension 2000, 35, 832–837. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Navedo, M.F.; Takeda, Y.; Nieves-Cintron, M.; Molkentin, J.D.; Santana, L.F. Elevated Ca2+ sparklet activity during acute hyperglycemia and diabetes in cerebral arterial smooth muscle cells. Am. J. Physiol. Cell. Physiol. 2010, 298, C211–C220. [Google Scholar] [CrossRef] [Green Version]
  206. Youm, J.B.; Park, K.S.; Jang, Y.J.; Leem, C.H. Effects of streptozotocin and unilateral nephrectomy on L-type Ca2+ channels and membrane capacitance in arteriolar smooth muscle cells. Pflugers Arch. 2015, 467, 1689–1697. [Google Scholar] [CrossRef]
  207. Blidner, A.G.; Rabinovich, G.A. ‘Sweetening’ pregnancy: Galectins at the fetomaternal interface. Am. J. Reprod. Immunol. 2013, 69, 369–382. [Google Scholar] [CrossRef]
  208. Tirado-Gonzalez, I.; Freitag, N.; Barrientos, G.; Shaikly, V.; Nagaeva, O.; Strand, M.; Kjellberg, L.; Klapp, B.F.; Mincheva-Nilsson, L.; Cohen, M.; et al. Galectin-1 influences trophoblast immune evasion and emerges as a predictive factor for the outcome of pregnancy. Mol. Hum. Reprod. 2013, 19, 43–53. [Google Scholar] [CrossRef] [Green Version]
  209. Chang, K.; Xiao, D.; Huang, X.; Xue, Z.; Yang, S.; Longo, L.D.; Zhang, L. Chronic hypoxia inhibits sex steroid hormone-mediated attenuation of ovine uterine arterial myogenic tone in pregnancy. Hypertension 2010, 56, 750–757. [Google Scholar] [CrossRef] [Green Version]
  210. Hu, X.Q.; Xiao, D.; Zhu, R.; Huang, X.; Yang, S.; Wilson, S.; Zhang, L. Pregnancy upregulates large-conductance Ca2+-activated K+ channel activity and attenuates myogenic tone in uterine arteries. Hypertension 2011, 58, 1132–1139. [Google Scholar] [CrossRef] [Green Version]
  211. Freitag, N.; Tirado-Gonzalez, I.; Barrientos, G.; Herse, F.; Thijssen, V.L.; Weedon-Fekjaer, S.M.; Schulz, H.; Wallukat, G.; Klapp, B.F.; Nevers, T.; et al. Interfering with Gal-1-mediated angiogenesis contributes to the pathogenesis of preeclampsia. Proc. Natl. Acad. Sci. USA 2013, 110, 11451–11456. [Google Scholar] [CrossRef] [Green Version]
  212. Jin, X.X.; Ying, X.; Dong, M.Y. Galectin-1 expression in the serum and placenta of pregnant women with fetal growth restriction and its significance. BMC Pregnancy Childbirth 2021, 21, 14. [Google Scholar] [CrossRef]
  213. Tejero, J.; Shiva, S.; Gladwin, M.T. Sources of Vascular Nitric Oxide and Reactive Oxygen Species and Their Regulation. Physiol. Rev. 2019, 99, 311–379. [Google Scholar] [CrossRef]
  214. Casas, A.I.; Nogales, C.; Mucke, H.A.M.; Petraina, A.; Cuadrado, A.; Rojo, A.I.; Ghezzi, P.; Jaquet, V.; Augsburger, F.; Dufrasne, F.; et al. On the Clinical Pharmacology of Reactive Oxygen Species. Pharmacol. Rev. 2020, 72, 801–828. [Google Scholar] [CrossRef]
  215. Checa, J.; Aran, J.M. Reactive Oxygen Species: Drivers of Physiological and Pathological Processes. J. Inflamm. Res. 2020, 13, 1057–1073. [Google Scholar] [CrossRef]
  216. Bedard, K.; Krause, K.H. The NOX family of ROS-generating NADPH oxidases: Physiology and pathophysiology. Physiol. Rev. 2007, 87, 245–313. [Google Scholar] [CrossRef]
  217. Montezano, A.C.; Burger, D.; Ceravolo, G.S.; Yusuf, H.; Montero, M.; Touyz, R.M. Novel Nox homologues in the vasculature: Focusing on Nox4 and Nox5. Clin. Sci. 2011, 120, 131–141. [Google Scholar] [CrossRef] [Green Version]
  218. Taylor, C.T. Mitochondria and cellular oxygen sensing in the HIF pathway. Biochem. J. 2008, 409, 19–26. [Google Scholar] [CrossRef] [Green Version]
  219. Turrens, J.F. Mitochondrial formation of reactive oxygen species. J. Physiol. 2003, 552, 335–344. [Google Scholar] [CrossRef]
  220. Murphy, M.P. How mitochondria produce reactive oxygen species. Biochem. J. 2009, 417, 1–13. [Google Scholar] [CrossRef] [Green Version]
  221. Rhee, S.G. Cell signaling. H2O2, a necessary evil for cell signaling. Science 2006, 312, 1882–1883. [Google Scholar] [CrossRef] [PubMed]
  222. Winterbourn, C.C. Hydrogen peroxide reactivity and specificity in thiol-based cell signalling. Biochem. Soc. Trans. 2020, 48, 745–754. [Google Scholar] [CrossRef] [PubMed]
  223. Pisoschi, A.M.; Pop, A.; Iordache, F.; Stanca, L.; Predoi, G.; Serban, A.I. Oxidative stress mitigation by antioxidants—An overview on their chemistry and influences on health status. Eur. J. Med. Chem. 2021, 209, 112891. [Google Scholar] [CrossRef] [PubMed]
  224. Kumar, K.V.; Das, U.N. Are free radicals involved in the pathobiology of human essential hypertension? Free Radic. Res. Commun. 1993, 19, 59–66. [Google Scholar] [CrossRef] [PubMed]
  225. Russo, C.; Olivieri, O.; Girelli, D.; Faccini, G.; Zenari, M.L.; Lombardi, S.; Corrocher, R. Anti-oxidant status and lipid peroxidation in patients with essential hypertension. J. Hypertens. 1998, 16, 1267–1271. [Google Scholar] [CrossRef] [PubMed]
  226. Koska, J.; Syrova, D.; Blazicek, P.; Marko, M.; Grna, J.D.; Kvetnansky, R.; Vigas, M. Malondialdehyde, lipofuscin and activity of antioxidant enzymes during physical exercise in patients with essential hypertension. J. Hypertens. 1999, 17, 529–535. [Google Scholar] [CrossRef]
  227. Redon, J.; Oliva, M.R.; Tormos, C.; Giner, V.; Chaves, J.; Iradi, A.; Saez, G.T. Antioxidant activities and oxidative stress byproducts in human hypertension. Hypertension 2003, 41, 1096–1101. [Google Scholar] [CrossRef] [Green Version]
  228. Simic, D.V.; Mimic-Oka, J.; Pljesa-Ercegovac, M.; Savic-Radojevic, A.; Opacic, M.; Matic, D.; Ivanovic, B.; Simic, T. Byproducts of oxidative protein damage and antioxidant enzyme activities in plasma of patients with different degrees of essential hypertension. J. Hum. Hypertens. 2006, 20, 149–155. [Google Scholar] [CrossRef]
  229. Rodrigo, R.; Prat, H.; Passalacqua, W.; Araya, J.; Bachler, J.P. Decrease in oxidative stress through supplementation of vitamins C and E is associated with a reduction in blood pressure in patients with essential hypertension. Clin. Sci. 2008, 114, 625–634. [Google Scholar] [CrossRef] [Green Version]
  230. Montezano, A.C.; Touyz, R.M. Reactive oxygen species, vascular Noxs, and hypertension: Focus on translational and clinical research. Antioxid. Redox Signal. 2014, 20, 164–182. [Google Scholar] [CrossRef] [Green Version]
  231. Minuz, P.; Patrignani, P.; Gaino, S.; Degan, M.; Menapace, L.; Tommasoli, R.; Seta, F.; Capone, M.L.; Tacconelli, S.; Palatresi, S.; et al. Increased oxidative stress and platelet activation in patients with hypertension and renovascular disease. Circulation 2002, 106, 2800–2805. [Google Scholar] [CrossRef] [Green Version]
  232. Will, J.C.; Ford, E.S.; Bowman, B.A. Serum vitamin C concentrations and diabetes: Findings from the Third National Health and Nutrition Examination Survey, 1988–1994. Am. J. Clin. Nutr. 1999, 70, 49–52. [Google Scholar] [CrossRef]
  233. De Cristofaro, R.; Rocca, B.; Vitacolonna, E.; Falco, A.; Marchesani, P.; Ciabattoni, G.; Landolfi, R.; Patrono, C.; Davi, G. Lipid and protein oxidation contribute to a prothrombotic state in patients with type 2 diabetes mellitus. J. Thromb. Haemost. 2003, 1, 250–256. [Google Scholar] [CrossRef]
  234. Palanduz, S.; Ademoglu, E.; Gokkusu, C.; Tamer, S. Plasma antioxidants and type 2 diabetes mellitus. Res. Commun. Mol. Pathol. Pharmacol. 2001, 109, 309–318. [Google Scholar]
  235. Komosinska-Vassev, K.; Olczyk, K.; Olczyk, P.; Winsz-Szczotka, K. Effects of metabolic control and vascular complications on indices of oxidative stress in type 2 diabetic patients. Diabetes Res. Clin. Pract. 2005, 68, 207–216. [Google Scholar] [CrossRef]
  236. Soliman, G.Z. Blood lipid peroxidation (superoxide dismutase, malondialdehyde, glutathione) levels in Egyptian type 2 diabetic patients. Singapore Med. J. 2008, 49, 129–136. [Google Scholar]
  237. Calabrese, V.; Cornelius, C.; Leso, V.; Trovato-Salinaro, A.; Ventimiglia, B.; Cavallaro, M.; Scuto, M.; Rizza, S.; Zanoli, L.; Neri, S.; et al. Oxidative stress, glutathione status, sirtuin and cellular stress response in type 2 diabetes. Biochim. Biophys. Acta 2012, 1822, 729–736. [Google Scholar] [CrossRef] [Green Version]
  238. Lutchmansingh, F.K.; Hsu, J.W.; Bennett, F.I.; Badaloo, A.V.; McFarlane-Anderson, N.; Gordon-Strachan, G.M.; Wright-Pascoe, R.A.; Jahoor, F.; Boyne, M.S. Glutathione metabolism in type 2 diabetes and its relationship with microvascular complications and glycemia. PLoS ONE 2018, 13, e0198626. [Google Scholar] [CrossRef] [Green Version]
  239. Huang, K.; Liang, Y.; Wang, K.; Wu, J.; Luo, H.; Yi, B. Influence of circulating nesfatin-1, GSH and SOD on insulin secretion in the development of T2DM. Front. Public Health 2022, 10, 882686. [Google Scholar] [CrossRef]
  240. Mikhail, M.S.; Anyaegbunam, A.; Garfinkel, D.; Palan, P.R.; Basu, J.; Romney, S.L. Preeclampsia and antioxidant nutrients: Decreased plasma levels of reduced ascorbic acid, alpha-tocopherol, and beta-carotene in women with preeclampsia. Am. J. Obstet. Gynecol. 1994, 171, 150–157. [Google Scholar] [CrossRef]
  241. Zusterzeel, P.L.; Mulder, T.P.; Peters, W.H.; Wiseman, S.A.; Steegers, E.A. Plasma protein carbonyls in nonpregnant, healthy pregnant and preeclamptic women. Free Radic. Res. 2000, 33, 471–476. [Google Scholar] [CrossRef] [PubMed]
  242. Madazli, R.; Benian, A.; Aydin, S.; Uzun, H.; Tolun, N. The plasma and placental levels of malondialdehyde, glutathione and superoxide dismutase in pre-eclampsia. J. Obstet. Gynaecol. 2002, 22, 477–480. [Google Scholar] [CrossRef] [PubMed]
  243. Aydin, S.; Benian, A.; Madazli, R.; Uludag, S.; Uzun, H.; Kaya, S. Plasma malondialdehyde, superoxide dismutase, sE-selectin, fibronectin, endothelin-1 and nitric oxide levels in women with preeclampsia. Eur. J. Obstet. Gynecol. Reprod. Biol. 2004, 113, 21–25. [Google Scholar] [CrossRef] [PubMed]
  244. Uzun, H.; Benian, A.; Madazli, R.; Topcuoglu, M.A.; Aydin, S.; Albayrak, M. Circulating oxidized low-density lipoprotein and paraoxonase activity in preeclampsia. Gynecol. Obstet. Investig. 2005, 60, 195–200. [Google Scholar] [CrossRef] [PubMed]
  245. Pimentel, A.M.; Pereira, N.R.; Costa, C.A.; Mann, G.E.; Cordeiro, V.S.; de Moura, R.S.; Brunini, T.M.; Mendes-Ribeiro, A.C.; Resende, A.C. L-arginine-nitric oxide pathway and oxidative stress in plasma and platelets of patients with pre-eclampsia. Hypertens. Res. 2013, 36, 783–788. [Google Scholar] [CrossRef] [Green Version]
  246. Kao, C.K.; Morton, J.S.; Quon, A.L.; Reyes, L.M.; Lopez-Jaramillo, P.; Davidge, S.T. Mechanism of vascular dysfunction due to circulating factors in women with pre-eclampsia. Clin. Sci. 2016, 130, 539–549. [Google Scholar] [CrossRef]
  247. Babic, G.M.; Markovic, S.D.; Varjacic, M.; Djordjevic, N.Z.; Nikolic, T.; Stojic, I.; Jakovljevic, V. Estradiol decreases blood pressure in association with redox regulation in preeclampsia. Clin. Exp. Hypertens. 2018, 40, 281–286. [Google Scholar] [CrossRef]
  248. Taravati, A.; Tohidi, F. Comprehensive analysis of oxidative stress markers and antioxidants status in preeclampsia. Taiwan J. Obstet. Gynecol. 2018, 57, 779–790. [Google Scholar] [CrossRef]
  249. Barden, A.; Beilin, L.J.; Ritchie, J.; Croft, K.D.; Walters, B.N.; Michael, C.A. Plasma and urinary 8-iso-prostane as an indicator of lipid peroxidation in pre-eclampsia and normal pregnancy. Clin. Sci. 1996, 91, 711–718. [Google Scholar] [CrossRef] [Green Version]
  250. Davi, G.; Ciabattoni, G.; Consoli, A.; Mezzetti, A.; Falco, A.; Santarone, S.; Pennese, E.; Vitacolonna, E.; Bucciarelli, T.; Costantini, F.; et al. In vivo formation of 8-iso-prostaglandin f2alpha and platelet activation in diabetes mellitus: Effects of improved metabolic control and vitamin E supplementation. Circulation 1999, 99, 224–229. [Google Scholar] [CrossRef] [Green Version]
  251. Swei, A.; Lacy, F.; DeLano, F.A.; Schmid-Schonbein, G.W. Oxidative stress in the Dahl hypertensive rat. Hypertension 1997, 30, 1628–1633. [Google Scholar] [CrossRef]
  252. Newaz, M.A.; Nawal, N.N. Effect of alpha-tocopherol on lipid peroxidation and total antioxidant status in spontaneously hypertensive rats. Am. J. Hypertens. 1998, 11, 1480–1485. [Google Scholar] [CrossRef] [Green Version]
  253. Schnackenberg, C.G.; Wilcox, C.S. Two-week administration of tempol attenuates both hypertension and renal excretion of 8-Iso prostaglandin f2alpha. Hypertension 1999, 33, 424–428. [Google Scholar] [CrossRef] [Green Version]
  254. Zhan, C.D.; Sindhu, R.K.; Pang, J.; Ehdaie, A.; Vaziri, N.D. Superoxide dismutase, catalase and glutathione peroxidase in the spontaneously hypertensive rat kidney: Effect of antioxidant-rich diet. J. Hypertens. 2004, 22, 2025–2033. [Google Scholar] [CrossRef]
  255. Viel, E.C.; Benkirane, K.; Javeshghani, D.; Touyz, R.M.; Schiffrin, E.L. Xanthine oxidase and mitochondria contribute to vascular superoxide anion generation in DOCA-salt hypertensive rats. Am. J. Physiol. Heart Circ. Physiol. 2008, 295, H281–H288. [Google Scholar] [CrossRef] [Green Version]
  256. Lerman, L.O.; Nath, K.A.; Rodriguez-Porcel, M.; Krier, J.D.; Schwartz, R.S.; Napoli, C.; Romero, J.C. Increased oxidative stress in experimental renovascular hypertension. Hypertension 2001, 37, 541–546. [Google Scholar] [CrossRef] [Green Version]
  257. Welch, W.J.; Mendonca, M.; Aslam, S.; Wilcox, C.S. Roles of oxidative stress and AT1 receptors in renal hemodynamics and oxygenation in the postclipped 2K,1C kidney. Hypertension 2003, 41, 692–696. [Google Scholar] [CrossRef] [Green Version]
  258. Guron, G.S.; Grimberg, E.S.; Basu, S.; Herlitz, H. Acute effects of the superoxide dismutase mimetic tempol on split kidney function in two-kidney one-clip hypertensive rats. J. Hypertens. 2006, 24, 387–394. [Google Scholar] [CrossRef]
  259. Vural, P.; Kabaca, G.; Firat, R.D.; Degirmencioglu, S. Administration of Selenium Decreases Lipid Peroxidation and Increases Vascular Endothelial Growth Factor in Streptozotocin Induced Diabetes Mellitus. Cell J. 2017, 19, 452–460. [Google Scholar] [CrossRef]
  260. Lasker, S.; Rahman, M.M.; Parvez, F.; Zamila, M.; Miah, P.; Nahar, K.; Kabir, F.; Sharmin, S.B.; Subhan, N.; Ahsan, G.U.; et al. High-fat diet-induced metabolic syndrome and oxidative stress in obese rats are ameliorated by yogurt supplementation. Sci. Rep. 2019, 9, 20026. [Google Scholar] [CrossRef] [Green Version]
  261. Qiu, Y.; Jiang, X.; Liu, D.; Deng, Z.; Hu, W.; Li, Z.; Li, Y. The Hypoglycemic and Renal Protection Properties of Crocin via Oxidative Stress-Regulated NF-kappaB Signaling in db/db Mice. Front. Pharmacol. 2020, 11, 541. [Google Scholar] [CrossRef] [PubMed]
  262. Gilani, S.J.; Bin-Jumah, M.N.; Al-Abbasi, F.A.; Nadeem, M.S.; Afzal, M.; Sayyed, N.; Kazmi, I. Fustin Ameliorates Elevated Levels of Leptin, Adiponectin, Serum TNF-alpha, and Intracellular Oxidative Free Radicals in High-Fat Diet and Streptozotocin-Induced Diabetic Rats. ACS Omega 2021, 6, 26098–26107. [Google Scholar] [CrossRef] [PubMed]
  263. Bauer, A.J.; Banek, C.T.; Needham, K.; Gillham, H.; Capoccia, S.; Regal, J.F.; Gilbert, J.S. Pravastatin attenuates hypertension, oxidative stress, and angiogenic imbalance in rat model of placental ischemia-induced hypertension. Hypertension 2013, 61, 1103–1110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Amaral, L.M.; Pinheiro, L.C.; Guimaraes, D.A.; Palei, A.C.; Sertorio, J.T.; Portella, R.L.; Tanus-Santos, J.E. Antihypertensive effects of inducible nitric oxide synthase inhibition in experimental pre-eclampsia. J. Cell. Mol. Med. 2013, 17, 1300–1307. [Google Scholar] [CrossRef] [PubMed]
  265. Badran, M.; Abuyassin, B.; Ayas, N.; Laher, I. Intermittent hypoxia impairs uterine artery function in pregnant mice. J. Physiol. 2019, 597, 2639–2650. [Google Scholar] [CrossRef] [PubMed]
  266. Rajagopalan, S.; Kurz, S.; Munzel, T.; Tarpey, M.; Freeman, B.A.; Griendling, K.K.; Harrison, D.G. Angiotensin II-mediated hypertension in the rat increases vascular superoxide production via membrane NADH/NADPH oxidase activation. Contribution to alterations of vasomotor tone. J. Clin. Investig. 1996, 97, 1916–1923. [Google Scholar] [CrossRef] [Green Version]
  267. Laursen, J.B.; Rajagopalan, S.; Galis, Z.; Tarpey, M.; Freeman, B.A.; Harrison, D.G. Role of superoxide in angiotensin II-induced but not catecholamine-induced hypertension. Circulation 1997, 95, 588–593. [Google Scholar] [CrossRef]
  268. Cifuentes, M.E.; Rey, F.E.; Carretero, O.A.; Pagano, P.J. Upregulation of p67(phox) and gp91(phox) in aortas from angiotensin II-infused mice. Am. J. Physiol. Heart Circ. Physiol. 2000, 279, H2234–H2240. [Google Scholar] [CrossRef] [Green Version]
  269. Beswick, R.A.; Zhang, H.; Marable, D.; Catravas, J.D.; Hill, W.D.; Webb, R.C. Long-term antioxidant administration attenuates mineralocorticoid hypertension and renal inflammatory response. Hypertension 2001, 37, 781–786. [Google Scholar] [CrossRef]
  270. Wu, R.; Millette, E.; Wu, L.; de Champlain, J. Enhanced superoxide anion formation in vascular tissues from spontaneously hypertensive and desoxycorticosterone acetate-salt hypertensive rats. J. Hypertens. 2001, 19, 741–748. [Google Scholar] [CrossRef]
  271. Landmesser, U.; Cai, H.; Dikalov, S.; McCann, L.; Hwang, J.; Jo, H.; Holland, S.M.; Harrison, D.G. Role of p47(phox) in vascular oxidative stress and hypertension caused by angiotensin II. Hypertension 2002, 40, 511–515. [Google Scholar] [CrossRef] [Green Version]
  272. Dantas, A.P.; Franco Mdo, C.; Silva-Antonialli, M.M.; Tostes, R.C.; Fortes, Z.B.; Nigro, D.; Carvalho, M.H. Gender differences in superoxide generation in microvessels of hypertensive rats: Role of NAD(P)H-oxidase. Cardiovasc. Res. 2004, 61, 22–29. [Google Scholar] [CrossRef] [Green Version]
  273. Zhang, Y.; Griendling, K.K.; Dikalova, A.; Owens, G.K.; Taylor, W.R. Vascular hypertrophy in angiotensin II-induced hypertension is mediated by vascular smooth muscle cell-derived H2O2. Hypertension 2005, 46, 732–737. [Google Scholar] [CrossRef] [Green Version]
  274. Dikalova, A.; Clempus, R.; Lassegue, B.; Cheng, G.; McCoy, J.; Dikalov, S.; San Martin, A.; Lyle, A.; Weber, D.S.; Weiss, D.; et al. Nox1 overexpression potentiates angiotensin II-induced hypertension and vascular smooth muscle hypertrophy in transgenic mice. Circulation 2005, 112, 2668–2676. [Google Scholar] [CrossRef]
  275. Zhou, X.; Bohlen, H.G.; Miller, S.J.; Unthank, J.L. NAD(P)H oxidase-derived peroxide mediates elevated basal and impaired flow-induced NO production in SHR mesenteric arteries in vivo. Am. J. Physiol. Heart Circ. Physiol. 2008, 295, H1008–H1016. [Google Scholar] [CrossRef] [Green Version]
  276. Tanito, M.; Nakamura, H.; Kwon, Y.W.; Teratani, A.; Masutani, H.; Shioji, K.; Kishimoto, C.; Ohira, A.; Horie, R.; Yodoi, J. Enhanced oxidative stress and impaired thioredoxin expression in spontaneously hypertensive rats. Antioxid. Redox Signal. 2004, 6, 89–97. [Google Scholar] [CrossRef]
  277. Chrissobolis, S.; Drummond, G.R.; Faraci, F.M.; Sobey, C.G. Chronic aldosterone administration causes Nox2-mediated increases in reactive oxygen species production and endothelial dysfunction in the cerebral circulation. J. Hypertens. 2014, 32, 1815–1821. [Google Scholar] [CrossRef] [Green Version]
  278. Rodrigues, D.; Costa, T.J.; Silva, J.F.; Neto, J.T.O.; Alves, J.V.; Fedoce, A.G.; Costa, R.M.; Tostes, R.C. Aldosterone Negatively Regulates Nrf2 Activity: An Additional Mechanism Contributing to Oxidative Stress and Vascular Dysfunction by Aldosterone. Int. J. Mol. Sci. 2021, 22, 6154. [Google Scholar] [CrossRef]
  279. Guzik, T.J.; Mussa, S.; Gastaldi, D.; Sadowski, J.; Ratnatunga, C.; Pillai, R.; Channon, K.M. Mechanisms of increased vascular superoxide production in human diabetes mellitus: Role of NAD(P)H oxidase and endothelial nitric oxide synthase. Circulation 2002, 105, 1656–1662. [Google Scholar] [CrossRef] [Green Version]
  280. San Martin, A.; Du, P.; Dikalova, A.; Lassegue, B.; Aleman, M.; Gongora, M.C.; Brown, K.; Joseph, G.; Harrison, D.G.; Taylor, W.R.; et al. Reactive oxygen species-selective regulation of aortic inflammatory gene expression in Type 2 diabetes. Am. J. Physiol. Heart Circ. Physiol. 2007, 292, H2073–H2082. [Google Scholar] [CrossRef]
  281. Erdei, N.; Bagi, Z.; Edes, I.; Kaley, G.; Koller, A. H2O2 increases production of constrictor prostaglandins in smooth muscle leading to enhanced arteriolar tone in Type 2 diabetic mice. Am. J. Physiol. Heart Circ. Physiol. 2007, 292, H649–H656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  282. Chettimada, S.; Ata, H.; Rawat, D.K.; Gulati, S.; Kahn, A.G.; Edwards, J.G.; Gupte, S.A. Contractile protein expression is upregulated by reactive oxygen species in aorta of Goto-Kakizaki rat. Am. J. Physiol. Heart Circ. Physiol. 2014, 306, H214–H224. [Google Scholar] [CrossRef] [Green Version]
  283. Cosentino, F.; Hishikawa, K.; Katusic, Z.S.; Luscher, T.F. High glucose increases nitric oxide synthase expression and superoxide anion generation in human aortic endothelial cells. Circulation 1997, 96, 25–28. [Google Scholar] [CrossRef] [PubMed]
  284. Inoguchi, T.; Li, P.; Umeda, F.; Yu, H.Y.; Kakimoto, M.; Imamura, M.; Aoki, T.; Etoh, T.; Hashimoto, T.; Naruse, M.; et al. High glucose level and free fatty acid stimulate reactive oxygen species production through protein kinase C--dependent activation of NAD(P)H oxidase in cultured vascular cells. Diabetes 2000, 49, 1939–1945. [Google Scholar] [CrossRef] [PubMed]
  285. Liu, S.; Ma, X.; Gong, M.; Shi, L.; Lincoln, T.; Wang, S. Glucose down-regulation of cGMP-dependent protein kinase I expression in vascular smooth muscle cells involves NAD(P)H oxidase-derived reactive oxygen species. Free Radic. Biol. Med. 2007, 42, 852–863. [Google Scholar] [CrossRef]
  286. Ding, H.; Aljofan, M.; Triggle, C.R. Oxidative stress and increased eNOS and NADPH oxidase expression in mouse microvessel endothelial cells. J. Cell. Physiol. 2007, 212, 682–689. [Google Scholar] [CrossRef]
  287. Xi, G.; Shen, X.; Maile, L.A.; Wai, C.; Gollahon, K.; Clemmons, D.R. Hyperglycemia enhances IGF-I-stimulated Src activation via increasing Nox4-derived reactive oxygen species in a PKCzeta-dependent manner in vascular smooth muscle cells. Diabetes 2012, 61, 104–113. [Google Scholar] [CrossRef] [Green Version]
  288. Many, A.; Hubel, C.A.; Fisher, S.J.; Roberts, J.M.; Zhou, Y. Invasive cytotrophoblasts manifest evidence of oxidative stress in preeclampsia. Am. J. Pathol. 2000, 156, 321–331. [Google Scholar] [CrossRef] [Green Version]
  289. Wang, Y.; Walsh, S.W. Increased superoxide generation is associated with decreased superoxide dismutase activity and mRNA expression in placental trophoblast cells in pre-eclampsia. Placenta 2001, 22, 206–212. [Google Scholar] [CrossRef]
  290. Sikkema, J.M.; van Rijn, B.B.; Franx, A.; Bruinse, H.W.; de Roos, R.; Stroes, E.S.; van Faassen, E.E. Placental superoxide is increased in pre-eclampsia. Placenta 2001, 22, 304–308. [Google Scholar] [CrossRef]
  291. Hnat, M.D.; Meadows, J.W.; Brockman, D.E.; Pitzer, B.; Lyall, F.; Myatt, L. Heat shock protein-70 and 4-hydroxy-2-nonenal adducts in human placental villous tissue of normotensive, preeclamptic and intrauterine growth restricted pregnancies. Am. J. Obstet. Gynecol. 2005, 193, 836–840. [Google Scholar] [CrossRef]
  292. Matsubara, K.; Matsubara, Y.; Hyodo, S.; Katayama, T.; Ito, M. Role of nitric oxide and reactive oxygen species in the pathogenesis of preeclampsia. J. Obstet. Gynaecol. Res. 2010, 36, 239–247. [Google Scholar] [CrossRef]
  293. Sedeek, M.; Gilbert, J.S.; LaMarca, B.B.; Sholook, M.; Chandler, D.L.; Wang, Y.; Granger, J.P. Role of reactive oxygen species in hypertension produced by reduced uterine perfusion in pregnant rats. Am. J. Hypertens. 2008, 21, 1152–1156. [Google Scholar] [CrossRef] [Green Version]
  294. Parrish, M.R.; Wallace, K.; Tam Tam, K.B.; Herse, F.; Weimer, A.; Wenzel, K.; Wallukat, G.; Ray, L.F.; Arany, M.; Cockrell, K.; et al. Hypertension in response to AT1-AA: Role of reactive oxygen species in pregnancy-induced hypertension. Am. J. Hypertens. 2011, 24, 835–840. [Google Scholar] [CrossRef]
  295. Richter, H.G.; Camm, E.J.; Modi, B.N.; Naeem, F.; Cross, C.M.; Cindrova-Davies, T.; Spasic-Boskovic, O.; Dunster, C.; Mudway, I.S.; Kelly, F.J.; et al. Ascorbate prevents placental oxidative stress and enhances birth weight in hypoxic pregnancy in rats. J. Physiol. 2012, 590, 1377–1387. [Google Scholar] [CrossRef]
  296. Morton, J.S.; Abdalvand, A.; Jiang, Y.; Sawamura, T.; Uwiera, R.R.; Davidge, S.T. Lectin-like oxidized low-density lipoprotein 1 receptor in a reduced uteroplacental perfusion pressure rat model of preeclampsia. Hypertension 2012, 59, 1014–1020. [Google Scholar] [CrossRef] [Green Version]
  297. Stanley, J.L.; Andersson, I.J.; Hirt, C.J.; Moore, L.; Dilworth, M.R.; Chade, A.R.; Sibley, C.P.; Davidge, S.T.; Baker, P.N. Effect of the anti-oxidant tempol on fetal growth in a mouse model of fetal growth restriction. Biol. Reprod. 2012, 87, 1–8. [Google Scholar] [CrossRef]
  298. Kusinski, L.C.; Stanley, J.L.; Dilworth, M.R.; Hirt, C.J.; Andersson, I.J.; Renshall, L.J.; Baker, B.C.; Baker, P.N.; Sibley, C.P.; Wareing, M.; et al. eNOS knockout mouse as a model of fetal growth restriction with an impaired uterine artery function and placental transport phenotype. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2012, 303, R86–R93. [Google Scholar] [CrossRef]
  299. Xiao, D.; Hu, X.Q.; Huang, X.; Zhou, J.; Wilson, S.M.; Yang, S.; Zhang, L. Chronic hypoxia during gestation enhances uterine arterial myogenic tone via heightened oxidative stress. PLoS ONE 2013, 8, e73731. [Google Scholar] [CrossRef] [Green Version]
  300. Rueda-Clausen, C.F.; Stanley, J.L.; Thambiraj, D.F.; Poudel, R.; Davidge, S.T.; Baker, P.N. Effect of prenatal hypoxia in transgenic mouse models of preeclampsia and fetal growth restriction. Reprod. Sci. 2014, 21, 492–502. [Google Scholar] [CrossRef] [Green Version]
  301. Ganguly, E.; Aljunaidy, M.M.; Kirschenman, R.; Spaans, F.; Morton, J.S.; Phillips, T.E.J.; Case, C.P.; Cooke, C.M.; Davidge, S.T. Sex-Specific Effects of Nanoparticle-Encapsulated MitoQ (nMitoQ) Delivery to the Placenta in a Rat Model of Fetal Hypoxia. Front. Physiol. 2019, 10, 562. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  302. Chen, G.; Lin, Y.; Chen, L.; Zeng, F.; Zhang, L.; Huang, Y.; Huang, P.; Liao, L.; Yu, Y. Role of DRAM1 in mitophagy contributes to preeclampsia regulation in mice. Mol. Med. Rep. 2020, 22, 1847–1858. [Google Scholar] [CrossRef] [PubMed]
  303. Morton, J.S.; Levasseur, J.; Ganguly, E.; Quon, A.; Kirschenman, R.; Dyck, J.R.B.; Fraser, G.M.; Davidge, S.T. Characterisation of the Selective Reduced Uteroplacental Perfusion (sRUPP) Model of Preeclampsia. Sci. Rep. 2019, 9, 9565. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Choi, S.; Kim, J.A.; Na, H.Y.; Kim, J.E.; Park, S.; Han, K.H.; Kim, Y.J.; Suh, S.H. NADPH oxidase 2-derived superoxide downregulates endothelial KCa3.1 in preeclampsia. Free Radic. Biol. Med. 2013, 57, 10–21. [Google Scholar] [CrossRef] [PubMed]
  305. Hu, X.Q.; Huang, X.; Xiao, D.; Zhang, L. Direct effect of chronic hypoxia in suppressing large conductance Ca2+-activated K+ channel activity in ovine uterine arteries via increasing oxidative stress. J. Physiol. 2016, 594, 343–356. [Google Scholar] [CrossRef] [PubMed]
  306. Wang, Y.; Walsh, S.W. Antioxidant activities and mRNA expression of superoxide dismutase, catalase, and glutathione peroxidase in normal and preeclamptic placentas. J. Soc. Gynecol. Investig. 1996, 3, 179–184. [Google Scholar] [CrossRef] [PubMed]
  307. Choi, H.; Allahdadi, K.J.; Tostes, R.C.; Webb, R.C. Augmented S-nitrosylation contributes to impaired relaxation in angiotensin II hypertensive mouse aorta: Role of thioredoxin reductase. J. Hypertens. 2011, 29, 2359–2368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  308. Lu, T.; Chai, Q.; Yu, L.; d’Uscio, L.V.; Katusic, Z.S.; He, T.; Lee, H.C. Reactive oxygen species signaling facilitates FOXO-3a/FBXO-dependent vascular BK channel beta1 subunit degradation in diabetic mice. Diabetes 2012, 61, 1860–1868. [Google Scholar] [CrossRef] [Green Version]
  309. Holland, O.J.; Cuffe, J.S.M.; Dekker Nitert, M.; Callaway, L.; Kwan Cheung, K.A.; Radenkovic, F.; Perkins, A.V. Placental mitochondrial adaptations in preeclampsia associated with progression to term delivery. Cell Death Dis. 2018, 9, 1150. [Google Scholar] [CrossRef] [Green Version]
  310. San Jose, G.; Moreno, M.U.; Olivan, S.; Beloqui, O.; Fortuno, A.; Diez, J.; Zalba, G. Functional effect of the p22phox -930A/G polymorphism on p22phox expression and NADPH oxidase activity in hypertension. Hypertension 2004, 44, 163–169. [Google Scholar] [CrossRef] [Green Version]
  311. Zalba, G.; Beaumont, F.J.; San Jose, G.; Fortuno, A.; Fortuno, M.A.; Etayo, J.C.; Diez, J. Vascular NADH/NADPH oxidase is involved in enhanced superoxide production in spontaneously hypertensive rats. Hypertension 2000, 35, 1055–1061. [Google Scholar] [CrossRef] [Green Version]
  312. Paravicini, T.M.; Chrissobolis, S.; Drummond, G.R.; Sobey, C.G. Increased NADPH-oxidase activity and Nox4 expression during chronic hypertension is associated with enhanced cerebral vasodilatation to NADPH in vivo. Stroke 2004, 35, 584–589. [Google Scholar] [CrossRef] [Green Version]
  313. Lodi, F.; Cogolludo, A.; Duarte, J.; Moreno, L.; Coviello, A.; Peral De Bruno, M.; Vera, R.; Galisteo, M.; Jimenez, R.; Tamargo, J.; et al. Increased NADPH oxidase activity mediates spontaneous aortic tone in genetically hypertensive rats. Eur. J. Pharmacol. 2006, 544, 97–103. [Google Scholar] [CrossRef]
  314. Akasaki, T.; Ohya, Y.; Kuroda, J.; Eto, K.; Abe, I.; Sumimoto, H.; Iida, M. Increased expression of gp91phox homologues of NAD(P)H oxidase in the aortic media during chronic hypertension: Involvement of the renin-angiotensin system. Hypertens. Res. 2006, 29, 813–820. [Google Scholar] [CrossRef]
  315. Wind, S.; Beuerlein, K.; Armitage, M.E.; Taye, A.; Kumar, A.H.; Janowitz, D.; Neff, C.; Shah, A.M.; Wingler, K.; Schmidt, H.H. Oxidative stress and endothelial dysfunction in aortas of aged spontaneously hypertensive rats by NOX1/2 is reversed by NADPH oxidase inhibition. Hypertension 2010, 56, 490–497. [Google Scholar] [CrossRef] [Green Version]
  316. Camargo, L.L.; Harvey, A.P.; Rios, F.J.; Tsiropoulou, S.; Da Silva, R.N.O.; Cao, Z.; Graham, D.; McMaster, C.; Burchmore, R.J.; Hartley, R.C.; et al. Vascular Nox (NADPH Oxidase) Compartmentalization, Protein Hyperoxidation, and Endoplasmic Reticulum Stress Response in Hypertension. Hypertension 2018, 72, 235–246. [Google Scholar] [CrossRef] [Green Version]
  317. Beswick, R.A.; Dorrance, A.M.; Leite, R.; Webb, R.C. NADH/NADPH oxidase and enhanced superoxide production in the mineralocorticoid hypertensive rat. Hypertension 2001, 38, 1107–1111. [Google Scholar] [CrossRef] [Green Version]
  318. Callera, G.E.; Touyz, R.M.; Tostes, R.C.; Yogi, A.; He, Y.; Malkinson, S.; Schiffrin, E.L. Aldosterone activates vascular p38MAP kinase and NADPH oxidase via c-Src. Hypertension 2005, 45, 773–779. [Google Scholar] [CrossRef] [Green Version]
  319. Miyata, K.; Rahman, M.; Shokoji, T.; Nagai, Y.; Zhang, G.X.; Sun, G.P.; Kimura, S.; Yukimura, T.; Kiyomoto, H.; Kohno, M.; et al. Aldosterone stimulates reactive oxygen species production through activation of NADPH oxidase in rat mesangial cells. J. Am. Soc. Nephrol. 2005, 16, 2906–2912. [Google Scholar] [CrossRef] [Green Version]
  320. Siuda, D.; Tobias, S.; Rus, A.; Xia, N.; Forstermann, U.; Li, H. Dexamethasone upregulates Nox1 expression in vascular smooth muscle cells. Pharmacology 2014, 94, 13–20. [Google Scholar] [CrossRef]
  321. Gayen, J.R.; Zhang, K.; RamachandraRao, S.P.; Mahata, M.; Chen, Y.; Kim, H.S.; Naviaux, R.K.; Sharma, K.; Mahata, S.K.; O’Connor, D.T. Role of reactive oxygen species in hyperadrenergic hypertension: Biochemical, physiological, and pharmacological evidence from targeted ablation of the chromogranin a (Chga) gene. Circ. Cardiovasc. Genet. 2010, 3, 414–425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  322. Fukui, T.; Ishizaka, N.; Rajagopalan, S.; Laursen, J.B.; Capers, Q.t.; Taylor, W.R.; Harrison, D.G.; de Leon, H.; Wilcox, J.N.; Griendling, K.K. p22phox mRNA expression and NADPH oxidase activity are increased in aortas from hypertensive rats. Circ. Res. 1997, 80, 45–51. [Google Scholar] [CrossRef] [PubMed]
  323. Mollnau, H.; Wendt, M.; Szocs, K.; Lassegue, B.; Schulz, E.; Oelze, M.; Li, H.; Bodenschatz, M.; August, M.; Kleschyov, A.L.; et al. Effects of angiotensin II infusion on the expression and function of NAD(P)H oxidase and components of nitric oxide/cGMP signaling. Circ. Res. 2002, 90, E58–E65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  324. Virdis, A.; Neves, M.F.; Amiri, F.; Touyz, R.M.; Schiffrin, E.L. Role of NAD(P)H oxidase on vascular alterations in angiotensin II-infused mice. J. Hypertens. 2004, 22, 535–542. [Google Scholar] [CrossRef] [PubMed]
  325. Matsuno, K.; Yamada, H.; Iwata, K.; Jin, D.; Katsuyama, M.; Matsuki, M.; Takai, S.; Yamanishi, K.; Miyazaki, M.; Matsubara, H.; et al. Nox1 is involved in angiotensin II-mediated hypertension: A study in Nox1-deficient mice. Circulation 2005, 112, 2677–2685. [Google Scholar] [CrossRef]
  326. Wang, D.; Chabrashvili, T.; Borrego, L.; Aslam, S.; Umans, J.G. Angiotensin II infusion alters vascular function in mouse resistance vessels: Roles of O and endothelium. J. Vasc. Res. 2006, 43, 109–119. [Google Scholar] [CrossRef]
  327. Oelze, M.; Daiber, A.; Brandes, R.P.; Hortmann, M.; Wenzel, P.; Hink, U.; Schulz, E.; Mollnau, H.; von Sandersleben, A.; Kleschyov, A.L.; et al. Nebivolol inhibits superoxide formation by NADPH oxidase and endothelial dysfunction in angiotensin II-treated rats. Hypertension 2006, 48, 677–684. [Google Scholar] [CrossRef] [Green Version]
  328. Touyz, R.M.; Chen, X.; Tabet, F.; Yao, G.; He, G.; Quinn, M.T.; Pagano, P.J.; Schiffrin, E.L. Expression of a functionally active gp91phox-containing neutrophil-type NAD(P)H oxidase in smooth muscle cells from human resistance arteries: Regulation by angiotensin II. Circ. Res. 2002, 90, 1205–1213. [Google Scholar] [CrossRef] [Green Version]
  329. Wendt, M.C.; Daiber, A.; Kleschyov, A.L.; Mulsch, A.; Sydow, K.; Schulz, E.; Chen, K.; Keaney, J.F., Jr.; Lassegue, B.; Walter, U.; et al. Differential effects of diabetes on the expression of the gp91phox homologues nox1 and nox4. Free Radic. Biol. Med. 2005, 39, 381–391. [Google Scholar] [CrossRef]
  330. Ding, H.; Hashem, M.; Triggle, C. Increased oxidative stress in the streptozotocin-induced diabetic apoE-deficient mouse: Changes in expression of NADPH oxidase subunits and eNOS. Eur. J. Pharmacol. 2007, 561, 121–128. [Google Scholar] [CrossRef]
  331. Wenzel, P.; Schulz, E.; Oelze, M.; Muller, J.; Schuhmacher, S.; Alhamdani, M.S.; Debrezion, J.; Hortmann, M.; Reifenberg, K.; Fleming, I.; et al. AT1-receptor blockade by telmisartan upregulates GTP-cyclohydrolase I and protects eNOS in diabetic rats. Free Radic. Biol. Med. 2008, 45, 619–626. [Google Scholar] [CrossRef]
  332. Rezende, F.; Moll, F.; Walter, M.; Helfinger, V.; Hahner, F.; Janetzko, P.; Ringel, C.; Weigert, A.; Fleming, I.; Weissmann, N.; et al. The NADPH organizers NoxO1 and p47phox are both mediators of diabetes-induced vascular dysfunction in mice. Redox Biol. 2018, 15, 12–21. [Google Scholar] [CrossRef]
  333. Manea, S.A.; Antonescu, M.L.; Fenyo, I.M.; Raicu, M.; Simionescu, M.; Manea, A. Epigenetic regulation of vascular NADPH oxidase expression and reactive oxygen species production by histone deacetylase-dependent mechanisms in experimental diabetes. Redox Biol. 2018, 16, 332–343. [Google Scholar] [CrossRef]
  334. Kassan, M.; Choi, S.K.; Galan, M.; Lee, Y.H.; Trebak, M.; Matrougui, K. Enhanced p22phox expression impairs vascular function through p38 and ERK1/2 MAP kinase-dependent mechanisms in type 2 diabetic mice. Am. J. Physiol. Heart Circ. Physiol. 2014, 306, H972–H980. [Google Scholar] [CrossRef]
  335. Bruder-Nascimento, T.; Callera, G.E.; Montezano, A.C.; He, Y.; Antunes, T.T.; Nguyen Dinh Cat, A.; Tostes, R.C.; Touyz, R.M. Vascular injury in diabetic db/db mice is ameliorated by atorvastatin: Role of Rac1/2-sensitive Nox-dependent pathways. Clin. Sci. 2015, 128, 411–423. [Google Scholar] [CrossRef]
  336. Quagliaro, L.; Piconi, L.; Assaloni, R.; Martinelli, L.; Motz, E.; Ceriello, A. Intermittent high glucose enhances apoptosis related to oxidative stress in human umbilical vein endothelial cells: The role of protein kinase C and NAD(P)H-oxidase activation. Diabetes 2003, 52, 2795–2804. [Google Scholar] [CrossRef] [Green Version]
  337. Weidig, P.; McMaster, D.; Bayraktutan, U. High glucose mediates pro-oxidant and antioxidant enzyme activities in coronary endothelial cells. Diabetes Obes. Metab. 2004, 6, 432–441. [Google Scholar] [CrossRef]
  338. Ulker, S.; McMaster, D.; McKeown, P.P.; Bayraktutan, U. Antioxidant vitamins C and E ameliorate hyperglycaemia-induced oxidative stress in coronary endothelial cells. Diabetes Obes. Metab. 2004, 6, 442–451. [Google Scholar] [CrossRef]
  339. Cui, X.L.; Brockman, D.; Campos, B.; Myatt, L. Expression of NADPH oxidase isoform 1 (Nox1) in human placenta: Involvement in preeclampsia. Placenta 2006, 27, 422–431. [Google Scholar] [CrossRef] [Green Version]
  340. Lim, R.; Acharya, R.; Delpachitra, P.; Hobson, S.; Sobey, C.G.; Drummond, G.R.; Wallace, E.M. Activin and NADPH-oxidase in preeclampsia: Insights from in vitro and murine studies. Am. J. Obstet. Gynecol. 2015, 212, 86.e1–86.e12. [Google Scholar] [CrossRef]
  341. Maynard, S.E.; Min, J.Y.; Merchan, J.; Lim, K.H.; Li, J.; Mondal, S.; Libermann, T.A.; Morgan, J.P.; Sellke, F.W.; Stillman, I.E.; et al. Excess placental soluble fms-like tyrosine kinase 1 (sFlt1) may contribute to endothelial dysfunction, hypertension, and proteinuria in preeclampsia. J. Clin. Investig. 2003, 111, 649–658. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  342. Peracoli, J.C.; Rudge, M.V.; Peracoli, M.T. Tumor necrosis factor-alpha in gestation and puerperium of women with gestational hypertension and pre-eclampsia. Am. J. Reprod. Immunol. 2007, 57, 177–185. [Google Scholar] [CrossRef] [PubMed]
  343. Siddiqui, A.H.; Irani, R.A.; Blackwell, S.C.; Ramin, S.M.; Kellems, R.E.; Xia, Y. Angiotensin receptor agonistic autoantibody is highly prevalent in preeclampsia: Correlation with disease severity. Hypertension 2010, 55, 386–393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  344. Rieber-Mohn, A.B.; Sugulle, M.; Wallukat, G.; Alnaes-Katjavivi, P.; Leite Storvold, G.; Bolstad, N.; Redman, C.W.; Dechend, R.; Staff, A.C. Auto-antibodies against the angiotensin II type I receptor in women with uteroplacental acute atherosis and preeclampsia at delivery and several years postpartum. J. Reprod. Immunol. 2018, 128, 23–29. [Google Scholar] [CrossRef] [PubMed]
  345. Sankaralingam, S.; Xu, Y.; Sawamura, T.; Davidge, S.T. Increased lectin-like oxidized low-density lipoprotein receptor-1 expression in the maternal vasculature of women with preeclampsia: Role for peroxynitrite. Hypertension 2009, 53, 270–277. [Google Scholar] [CrossRef] [Green Version]
  346. Tam Tam, K.B.; Lamarca, B.; Arany, M.; Cockrell, K.; Fournier, L.; Murphy, S.; Martin, J.N., Jr.; Granger, J.P. Role of reactive oxygen species during hypertension in response to chronic antiangiogenic factor (sFlt-1) excess in pregnant rats. Am. J. Hypertens. 2011, 24, 110–113. [Google Scholar] [CrossRef] [Green Version]
  347. Brewer, J.; Liu, R.; Lu, Y.; Scott, J.; Wallace, K.; Wallukat, G.; Moseley, J.; Herse, F.; Dechend, R.; Martin, J.N., Jr.; et al. Endothelin-1, oxidative stress, and endogenous angiotensin II: Mechanisms of angiotensin II type I receptor autoantibody-enhanced renal and blood pressure response during pregnancy. Hypertension 2013, 62, 886–892. [Google Scholar] [CrossRef]
  348. Santana-Garrido, A.; Reyes-Goya, C.; Espinosa-Martin, P.; Sobrevia, L.; Beltran, L.M.; Vazquez, C.M.; Mate, A. Oxidative and Inflammatory Imbalance in Placenta and Kidney of sFlt1-Induced Early-Onset Preeclampsia Rat Model. Antioxidants 2022, 11, 1608. [Google Scholar] [CrossRef]
  349. Muttukrishna, S.; Knight, P.G.; Groome, N.P.; Redman, C.W.; Ledger, W.L. Activin A and inhibin A as possible endocrine markers for pre-eclampsia. Lancet 1997, 349, 1285–1288. [Google Scholar] [CrossRef]
  350. Dechend, R.; Viedt, C.; Muller, D.N.; Ugele, B.; Brandes, R.P.; Wallukat, G.; Park, J.K.; Janke, J.; Barta, P.; Theuer, J.; et al. AT1 receptor agonistic antibodies from preeclamptic patients stimulate NADPH oxidase. Circulation 2003, 107, 1632–1639. [Google Scholar] [CrossRef] [Green Version]
  351. Choi, S.; Kim, J.A.; Li, H.Y.; Lee, S.J.; Seok, Y.S.; Kim, T.H.; Han, K.H.; Park, M.H.; Cho, G.J.; Suh, S.H. Altered Redox State Modulates Endothelial KCa2.3 and KCa3.1 Levels in Normal Pregnancy and Preeclampsia. Antioxid. Redox Signal. 2019, 30, 505–519. [Google Scholar] [CrossRef]
  352. De Keulenaer, G.W.; Alexander, R.W.; Ushio-Fukai, M.; Ishizaka, N.; Griendling, K.K. Tumour necrosis factor alpha activates a p22phox-based NADH oxidase in vascular smooth muscle. Biochem. J. 1998, 329 Pt 3, 653–657. [Google Scholar] [CrossRef]
  353. Sun, Y.; Chen, X. Ox-LDL-induced LOX-1 expression in vascular smooth muscle cells: Role of reactive oxygen species. Fundam. Clin. Pharmacol. 2011, 25, 572–579. [Google Scholar] [CrossRef]
  354. Chen, X.P.; Xun, K.L.; Wu, Q.; Zhang, T.T.; Shi, J.S.; Du, G.H. Oxidized low density lipoprotein receptor-1 mediates oxidized low density lipoprotein-induced apoptosis in human umbilical vein endothelial cells: Role of reactive oxygen species. Vascul. Pharmacol. 2007, 47, 1–9. [Google Scholar] [CrossRef]
  355. Tao, R.; Coleman, M.C.; Pennington, J.D.; Ozden, O.; Park, S.H.; Jiang, H.; Kim, H.S.; Flynn, C.R.; Hill, S.; Hayes McDonald, W.; et al. Sirt3-mediated deacetylation of evolutionarily conserved lysine 122 regulates MnSOD activity in response to stress. Mol. Cell 2010, 40, 893–904. [Google Scholar] [CrossRef] [Green Version]
  356. Chen, Y.; Zhang, J.; Lin, Y.; Lei, Q.; Guan, K.L.; Zhao, S.; Xiong, Y. Tumour suppressor SIRT3 deacetylates and activates manganese superoxide dismutase to scavenge ROS. EMBO Rep. 2011, 12, 534–541. [Google Scholar] [CrossRef]
  357. Dikalova, A.E.; Pandey, A.; Xiao, L.; Arslanbaeva, L.; Sidorova, T.; Lopez, M.G.; Billings, F.T.t.; Verdin, E.; Auwerx, J.; Harrison, D.G.; et al. Mitochondrial Deacetylase Sirt3 Reduces Vascular Dysfunction and Hypertension While Sirt3 Depletion in Essential Hypertension Is Linked to Vascular Inflammation and Oxidative Stress. Circ. Res. 2020, 126, 439–452. [Google Scholar] [CrossRef] [Green Version]
  358. De Cavanagh, E.M.; Toblli, J.E.; Ferder, L.; Piotrkowski, B.; Stella, I.; Fraga, C.G.; Inserra, F. Angiotensin II blockade improves mitochondrial function in spontaneously hypertensive rats. Cell. Mol. Biol. 2005, 51, 573–578. [Google Scholar]
  359. Rodriguez-Iturbe, B.; Sepassi, L.; Quiroz, Y.; Ni, Z.; Wallace, D.C.; Vaziri, N.D. Association of mitochondrial SOD deficiency with salt-sensitive hypertension and accelerated renal senescence. J. Appl. Physiol. 2007, 102, 255–260. [Google Scholar] [CrossRef] [Green Version]
  360. Dikalova, A.E.; Itani, H.A.; Nazarewicz, R.R.; McMaster, W.G.; Flynn, C.R.; Uzhachenko, R.; Fessel, J.P.; Gamboa, J.L.; Harrison, D.G.; Dikalov, S.I. Sirt3 Impairment and SOD2 Hyperacetylation in Vascular Oxidative Stress and Hypertension. Circ. Res. 2017, 121, 564–574. [Google Scholar] [CrossRef]
  361. Porter, G.A., Jr.; Beutner, G. Cyclophilin D, Somehow a Master Regulator of Mitochondrial Function. Biomolecules 2018, 8, 176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  362. Itani, H.A.; Dikalova, A.E.; McMaster, W.G.; Nazarewicz, R.R.; Bikineyeva, A.T.; Harrison, D.G.; Dikalov, S.I. Mitochondrial Cyclophilin D in Vascular Oxidative Stress and Hypertension. Hypertension 2016, 67, 1218–1227. [Google Scholar] [CrossRef] [PubMed]
  363. Peng, S.Y.; Tsai, C.H.; Wu, X.M.; Huang, H.H.; Chen, Z.W.; Lee, B.C.; Chang, Y.Y.; Pan, C.T.; Wu, V.C.; Chou, C.H.; et al. Aldosterone Suppresses Endothelial Mitochondria through Mineralocorticoid Receptor/Mitochondrial Reactive Oxygen Species Pathway. Biomedicines 2022, 10, 1119. [Google Scholar] [CrossRef] [PubMed]
  364. Zorov, D.B.; Juhaszova, M.; Sollott, S.J. Mitochondrial reactive oxygen species (ROS) and ROS-induced ROS release. Physiol. Rev. 2014, 94, 909–950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  365. Dikalov, S.I.; Nazarewicz, R.R.; Bikineyeva, A.; Hilenski, L.; Lassegue, B.; Griendling, K.K.; Harrison, D.G.; Dikalova, A.E. Nox2-induced production of mitochondrial superoxide in angiotensin II-mediated endothelial oxidative stress and hypertension. Antioxid. Redox Signal. 2014, 20, 281–294. [Google Scholar] [CrossRef]
  366. Kizhakekuttu, T.J.; Wang, J.; Dharmashankar, K.; Ying, R.; Gutterman, D.D.; Vita, J.A.; Widlansky, M.E. Adverse alterations in mitochondrial function contribute to type 2 diabetes mellitus-related endothelial dysfunction in humans. Arterioscler. Thromb. Vasc. Biol. 2012, 32, 2531–2539. [Google Scholar] [CrossRef] [Green Version]
  367. Mackenzie, R.M.; Salt, I.P.; Miller, W.H.; Logan, A.; Ibrahim, H.A.; Degasperi, A.; Dymott, J.A.; Hamilton, C.A.; Murphy, M.P.; Delles, C.; et al. Mitochondrial reactive oxygen species enhance AMP-activated protein kinase activation in the endothelium of patients with coronary artery disease and diabetes. Clin. Sci. 2013, 124, 403–411. [Google Scholar] [CrossRef] [Green Version]
  368. Makino, A.; Scott, B.T.; Dillmann, W.H. Mitochondrial fragmentation and superoxide anion production in coronary endothelial cells from a mouse model of type 1 diabetes. Diabetologia 2010, 53, 1783–1794. [Google Scholar] [CrossRef] [Green Version]
  369. Cho, Y.E.; Basu, A.; Dai, A.; Heldak, M.; Makino, A. Coronary endothelial dysfunction and mitochondrial reactive oxygen species in type 2 diabetic mice. Am. J. Physiol. Cell. Physiol. 2013, 305, C1033–C1040. [Google Scholar] [CrossRef] [Green Version]
  370. Keller, A.C.; Knaub, L.A.; McClatchey, P.M.; Connon, C.A.; Bouchard, R.; Miller, M.W.; Geary, K.E.; Walker, L.A.; Klemm, D.J.; Reusch, J.E. Differential Mitochondrial Adaptation in Primary Vascular Smooth Muscle Cells from a Diabetic Rat Model. Oxid. Med. Cell. Longev. 2016, 2016, 8524267. [Google Scholar] [CrossRef] [Green Version]
  371. Merdzo, I.; Rutkai, I.; Sure, V.N.; McNulty, C.A.; Katakam, P.V.; Busija, D.W. Impaired Mitochondrial Respiration in Large Cerebral Arteries of Rats with Type 2 Diabetes. J. Vasc. Res. 2017, 54, 1–12. [Google Scholar] [CrossRef] [Green Version]
  372. Merdzo, I.; Rutkai, I.; Sure, V.; Katakam, P.V.G.; Busija, D.W. Effects of prolonged type 2 diabetes on mitochondrial function in cerebral blood vessels. Am. J. Physiol. Heart Circ. Physiol. 2019, 317, H1086–H1092. [Google Scholar] [CrossRef]
  373. Nishikawa, T.; Edelstein, D.; Du, X.L.; Yamagishi, S.; Matsumura, T.; Kaneda, Y.; Yorek, M.A.; Beebe, D.; Oates, P.J.; Hammes, H.P.; et al. Normalizing mitochondrial superoxide production blocks three pathways of hyperglycaemic damage. Nature 2000, 404, 787–790. [Google Scholar] [CrossRef]
  374. Quagliaro, L.; Piconi, L.; Assaloni, R.; Da Ros, R.; Maier, A.; Zuodar, G.; Ceriello, A. Intermittent high glucose enhances ICAM-1, VCAM-1 and E-selectin expression in human umbilical vein endothelial cells in culture: The distinct role of protein kinase C and mitochondrial superoxide production. Atherosclerosis 2005, 183, 259–267. [Google Scholar] [CrossRef]
  375. Yu, T.; Sheu, S.S.; Robotham, J.L.; Yoon, Y. Mitochondrial fission mediates high glucose-induced cell death through elevated production of reactive oxygen species. Cardiovasc. Res. 2008, 79, 341–351. [Google Scholar] [CrossRef]
  376. Xie, L.; Zhu, X.; Hu, Y.; Li, T.; Gao, Y.; Shi, Y.; Tang, S. Mitochondrial DNA oxidative damage triggering mitochondrial dysfunction and apoptosis in high glucose-induced HRECs. Investig. Ophthalmol. Vis. Sci. 2008, 49, 4203–4209. [Google Scholar] [CrossRef]
  377. Ungvari, Z.; Labinskyy, N.; Mukhopadhyay, P.; Pinto, J.T.; Bagi, Z.; Ballabh, P.; Zhang, C.; Pacher, P.; Csiszar, A. Resveratrol attenuates mitochondrial oxidative stress in coronary arterial endothelial cells. Am. J. Physiol. Heart Circ. Physiol. 2009, 297, H1876–H1881. [Google Scholar] [CrossRef] [Green Version]
  378. Sun, J.; Xu, Y.; Sun, S.; Sun, Y.; Wang, X. Intermittent high glucose enhances cell proliferation and VEGF expression in retinal endothelial cells: The role of mitochondrial reactive oxygen species. Mol. Cell. Biochem. 2010, 343, 27–35. [Google Scholar] [CrossRef]
  379. Shenouda, S.M.; Widlansky, M.E.; Chen, K.; Xu, G.; Holbrook, M.; Tabit, C.E.; Hamburg, N.M.; Frame, A.A.; Caiano, T.L.; Kluge, M.A.; et al. Altered mitochondrial dynamics contributes to endothelial dysfunction in diabetes mellitus. Circulation 2011, 124, 444–453. [Google Scholar] [CrossRef] [Green Version]
  380. Ren, X.; Ren, L.; Wei, Q.; Shao, H.; Chen, L.; Liu, N. Advanced glycation end-products decreases expression of endothelial nitric oxide synthase through oxidative stress in human coronary artery endothelial cells. Cardiovasc. Diabetol. 2017, 16, 52. [Google Scholar] [CrossRef] [Green Version]
  381. Yoshinaga, A.; Kajihara, N.; Kukidome, D.; Motoshima, H.; Matsumura, T.; Nishikawa, T.; Araki, E. Hypoglycemia Induces Mitochondrial Reactive Oxygen Species Production Through Increased Fatty Acid Oxidation and Promotes Retinal Vascular Permeability in Diabetic Mice. Antioxid. Redox Signal. 2021, 34, 1245–1259. [Google Scholar] [CrossRef] [PubMed]
  382. Yung, H.W.; Colleoni, F.; Dommett, E.; Cindrova-Davies, T.; Kingdom, J.; Murray, A.J.; Burton, G.J. Noncanonical mitochondrial unfolded protein response impairs placental oxidative phosphorylation in early-onset preeclampsia. Proc. Natl. Acad. Sci. USA 2019, 116, 18109–18118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  383. Vangrieken, P.; Al-Nasiry, S.; Bast, A.; Leermakers, P.A.; Tulen, C.B.M.; Schiffers, P.M.H.; van Schooten, F.J.; Remels, A.H.V. Placental Mitochondrial Abnormalities in Preeclampsia. Reprod. Sci. 2021, 28, 2186–2199. [Google Scholar] [CrossRef] [PubMed]
  384. Wang, Y.; Walsh, S.W. Placental mitochondria as a source of oxidative stress in pre-eclampsia. Placenta 1998, 19, 581–586. [Google Scholar] [CrossRef] [PubMed]
  385. Shibata, E.; Nanri, H.; Ejima, K.; Araki, M.; Fukuda, J.; Yoshimura, K.; Toki, N.; Ikeda, M.; Kashimura, M. Enhancement of mitochondrial oxidative stress and up-regulation of antioxidant protein peroxiredoxin III/SP-22 in the mitochondria of human pre-eclamptic placentae. Placenta 2003, 24, 698–705. [Google Scholar] [CrossRef]
  386. Vaka, R.; Deer, E.; Cunningham, M.; McMaster, K.M.; Wallace, K.; Cornelius, D.C.; Amaral, L.M.; LaMarca, B. Characterization of Mitochondrial Bioenergetics in Preeclampsia. J. Clin. Med. 2021, 10, 5063. [Google Scholar] [CrossRef]
  387. Vaka, V.R.; McMaster, K.M.; Cunningham, M.W., Jr.; Ibrahim, T.; Hazlewood, R.; Usry, N.; Cornelius, D.C.; Amaral, L.M.; LaMarca, B. Role of Mitochondrial Dysfunction and Reactive Oxygen Species in Mediating Hypertension in the Reduced Uterine Perfusion Pressure Rat Model of Preeclampsia. Hypertension 2018, 72, 703–711. [Google Scholar] [CrossRef]
  388. Vaka, V.R.; Cunningham, M.W.; Deer, E.; Franks, M.; Ibrahim, T.; Amaral, L.M.; Usry, N.; Cornelius, D.C.; Dechend, R.; Wallukat, G.; et al. Blockade of endogenous angiotensin II type I receptor agonistic autoantibody activity improves mitochondrial reactive oxygen species and hypertension in a rat model of preeclampsia. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2020, 318, R256–R262. [Google Scholar] [CrossRef]
  389. Yang, Y.; Xu, P.; Zhu, F.; Liao, J.; Wu, Y.; Hu, M.; Fu, H.; Qiao, J.; Lin, L.; Huang, B.; et al. The Potent Antioxidant MitoQ Protects Against Preeclampsia During Late Gestation but Increases the Risk of Preeclampsia When Administered in Early Pregnancy. Antioxid. Redox Signal. 2021, 34, 118–136. [Google Scholar] [CrossRef]
  390. Jayaram, A.; Deer, E.; Amaral, L.M.; Campbell, N.; Vaka, V.R.; Cunningham, M.; Ibrahim, T.; Cornelius, D.C.; LaMarca, B.B. The role of tumor necrosis factor in triggering activation of natural killer cell, multi-organ mitochondrial dysfunction and hypertension during pregnancy. Pregnancy Hypertens. 2021, 24, 65–72. [Google Scholar] [CrossRef]
  391. McCarthy, C.; Kenny, L.C. Therapeutically targeting mitochondrial redox signalling alleviates endothelial dysfunction in preeclampsia. Sci. Rep. 2016, 6, 32683. [Google Scholar] [CrossRef] [Green Version]
  392. Deer, E.; Vaka, V.R.; McMaster, K.M.; Wallace, K.; Cornelius, D.C.; Amaral, L.M.; Cunningham, M.W.; LaMarca, B. Vascular endothelial mitochondrial oxidative stress in response to preeclampsia: A role for angiotension II type 1 autoantibodies. Am. J. Obstet. Gynecol. MFM 2021, 3, 100275. [Google Scholar] [CrossRef]
  393. Sanchez-Aranguren, L.C.; Espinosa-Gonzalez, C.T.; Gonzalez-Ortiz, L.M.; Sanabria-Barrera, S.M.; Riano-Medina, C.E.; Nunez, A.F.; Ahmed, A.; Vasquez-Vivar, J.; Lopez, M. Soluble Fms-Like Tyrosine Kinase-1 Alters Cellular Metabolism and Mitochondrial Bioenergetics in Preeclampsia. Front. Physiol. 2018, 9, 83. [Google Scholar] [CrossRef] [Green Version]
  394. Colleoni, F.; Padmanabhan, N.; Yung, H.W.; Watson, E.D.; Cetin, I.; Tissot van Patot, M.C.; Burton, G.J.; Murray, A.J. Suppression of mitochondrial electron transport chain function in the hypoxic human placenta: A role for miRNA-210 and protein synthesis inhibition. PLoS ONE 2013, 8, e55194. [Google Scholar] [CrossRef]
  395. Vangrieken, P.; Al-Nasiry, S.; Bast, A.; Leermakers, P.A.; Tulen, C.B.M.; Janssen, G.M.J.; Kaminski, I.; Geomini, I.; Lemmens, T.; Schiffers, P.M.H.; et al. Hypoxia-induced mitochondrial abnormalities in cells of the placenta. PLoS ONE 2021, 16, e0245155. [Google Scholar] [CrossRef]
  396. Hu, X.Q.; Song, R.; Dasgupta, C.; Romero, M.; Juarez, R.; Hanson, J.; Blood, A.B.; Wilson, S.M.; Zhang, L. MicroRNA-210-mediated mtROS confer hypoxia-induced suppression of STOCs in ovine uterine arteries. Br. J. Pharmacol. 2022, 179, 4640–4654. [Google Scholar] [CrossRef]
  397. Shah, M.S.; Brownlee, M. Molecular and Cellular Mechanisms of Cardiovascular Disorders in Diabetes. Circ. Res. 2016, 118, 1808–1829. [Google Scholar] [CrossRef] [Green Version]
  398. Togliatto, G.; Lombardo, G.; Brizzi, M.F. The Future Challenge of Reactive Oxygen Species (ROS) in Hypertension: From Bench to Bed Side. Int. J. Mol. Sci. 2017, 18, 1988. [Google Scholar] [CrossRef] [Green Version]
  399. Chiarello, D.I.; Abad, C.; Rojas, D.; Toledo, F.; Vazquez, C.M.; Mate, A.; Sobrevia, L.; Marin, R. Oxidative stress: Normal pregnancy versus preeclampsia. Biochim. Biophys. Acta Mol. Basis Dis. 2020, 1866, 165354. [Google Scholar] [CrossRef]
  400. Hu, X.Q.; Zhang, L. Hypoxia and Mitochondrial Dysfunction in Pregnancy Complications. Antioxidants 2021, 10, 405. [Google Scholar] [CrossRef]
  401. Hu, X.Q.; Zhang, L. Mitochondrial Dysfunction in the Pathogenesis of Preeclampsia. Curr. Hypertens. Rep. 2022, 24, 157–172. [Google Scholar] [CrossRef] [PubMed]
  402. Schnackenberg, C.G.; Welch, W.J.; Wilcox, C.S. Normalization of blood pressure and renal vascular resistance in SHR with a membrane-permeable superoxide dismutase mimetic: Role of nitric oxide. Hypertension 1998, 32, 59–64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  403. Park, J.B.; Touyz, R.M.; Chen, X.; Schiffrin, E.L. Chronic treatment with a superoxide dismutase mimetic prevents vascular remodeling and progression of hypertension in salt-loaded stroke-prone spontaneously hypertensive rats. Am. J. Hypertens. 2002, 15, 78–84. [Google Scholar] [CrossRef] [PubMed]
  404. Palm, F.; Onozato, M.; Welch, W.J.; Wilcox, C.S. Blood pressure, blood flow, and oxygenation in the clipped kidney of chronic 2-kidney, 1-clip rats: Effects of tempol and Angiotensin blockade. Hypertension 2010, 55, 298–304. [Google Scholar] [CrossRef] [PubMed]
  405. Brands, M.W.; Bell, T.D.; Gibson, B. Nitric oxide may prevent hypertension early in diabetes by counteracting renal actions of superoxide. Hypertension 2004, 43, 57–63. [Google Scholar] [CrossRef]
  406. Onuma, S.; Nakanishi, K. Superoxide dismustase mimetic tempol decreases blood pressure by increasing renal medullary blood flow in hyperinsulinemic-hypertensive rats. Metabolism 2004, 53, 1305–1308. [Google Scholar] [CrossRef]
  407. Peixoto, E.B.; Pessoa, B.S.; Biswas, S.K.; Lopes de Faria, J.B. Antioxidant SOD mimetic prevents NADPH oxidase-induced oxidative stress and renal damage in the early stage of experimental diabetes and hypertension. Am. J. Nephrol. 2009, 29, 309–318. [Google Scholar] [CrossRef]
  408. Li, H.; Liu, X.; Ren, Z.; Gu, J.; Lu, Y.; Wang, X.; Zhang, L. Effects of Diabetic Hyperglycemia on Central Ang-(1-7)-Mas-R-nNOS Pathways in Spontaneously Hypertensive Rats. Cell. Physiol. Biochem. 2016, 40, 1186–1197. [Google Scholar] [CrossRef]
  409. Hoffmann, D.S.; Weydert, C.J.; Lazartigues, E.; Kutschke, W.J.; Kienzle, M.F.; Leach, J.E.; Sharma, J.A.; Sharma, R.V.; Davisson, R.L. Chronic tempol prevents hypertension, proteinuria, and poor feto-placental outcomes in BPH/5 mouse model of preeclampsia. Hypertension 2008, 51, 1058–1065. [Google Scholar] [CrossRef] [Green Version]
  410. Carlstrom, M.; Brown, R.D.; Sallstrom, J.; Larsson, E.; Zilmer, M.; Zabihi, S.; Eriksson, U.J.; Persson, A.E. SOD1 deficiency causes salt sensitivity and aggravates hypertension in hydronephrosis. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2009, 297, R82–R92. [Google Scholar] [CrossRef] [Green Version]
  411. Baumer, A.T.; Kruger, C.A.; Falkenberg, J.; Freyhaus, H.T.; Rosen, R.; Fink, K.; Rosenkranz, S. The NAD(P)H oxidase inhibitor apocynin improves endothelial NO/superoxide balance and lowers effectively blood pressure in spontaneously hypertensive rats: Comparison to calcium channel blockade. Clin. Exp. Hypertens. 2007, 29, 287–299. [Google Scholar] [CrossRef]
  412. Unger, B.S.; Patil, B.M. Apocynin improves endothelial function and prevents the development of hypertension in fructose fed rat. Indian J. Pharmacol. 2009, 41, 208–212. [Google Scholar] [CrossRef]
  413. Guimaraes, D.D.; Carvalho, C.C.; Braga, V.A. Scavenging of NADPH oxidase-derived superoxide anions improves depressed baroreflex sensitivity in spontaneously hypertensive rats. Clin. Exp. Pharmacol. Physiol. 2012, 39, 373–378. [Google Scholar] [CrossRef]
  414. Wallace, K.; Cornelius, D.C.; Scott, J.; Heath, J.; Moseley, J.; Chatman, K.; LaMarca, B. CD4+ T cells are important mediators of oxidative stress that cause hypertension in response to placental ischemia. Hypertension 2014, 64, 1151–1158. [Google Scholar] [CrossRef] [Green Version]
  415. Perassa, L.A.; Graton, M.E.; Potje, S.R.; Troiano, J.A.; Lima, M.S.; Vale, G.T.; Pereira, A.A.; Nakamune, A.C.; Sumida, D.H.; Tirapelli, C.R.; et al. Apocynin reduces blood pressure and restores the proper function of vascular endothelium in SHR. Vasc. Pharmacol. 2016, 87, 38–48. [Google Scholar] [CrossRef]
  416. Dikalova, A.E.; Gongora, M.C.; Harrison, D.G.; Lambeth, J.D.; Dikalov, S.; Griendling, K.K. Upregulation of Nox1 in vascular smooth muscle leads to impaired endothelium-dependent relaxation via eNOS uncoupling. Am. J. Physiol. Heart Circ. Physiol. 2010, 299, H673–H679. [Google Scholar] [CrossRef] [Green Version]
  417. Youn, J.Y.; Gao, L.; Cai, H. The p47phox- and NADPH oxidase organiser 1 (NOXO1)-dependent activation of NADPH oxidase 1 (NOX1) mediates endothelial nitric oxide synthase (eNOS) uncoupling and endothelial dysfunction in a streptozotocin-induced murine model of diabetes. Diabetologia 2012, 55, 2069–2079. [Google Scholar] [CrossRef] [Green Version]
  418. Graham, D.; Huynh, N.N.; Hamilton, C.A.; Beattie, E.; Smith, R.A.; Cocheme, H.M.; Murphy, M.P.; Dominiczak, A.F. Mitochondria-targeted antioxidant MitoQ10 improves endothelial function and attenuates cardiac hypertrophy. Hypertension 2009, 54, 322–328. [Google Scholar] [CrossRef] [Green Version]
  419. Dikalova, A.; Mayorov, V.; Xiao, L.; Panov, A.; Amarnath, V.; Zagol-Ikapitte, I.; Vergeade, A.; Ao, M.; Yermalitsky, V.; Nazarewicz, R.R.; et al. Mitochondrial Isolevuglandins Contribute to Vascular Oxidative Stress and Mitochondria-Targeted Scavenger of Isolevuglandins Reduces Mitochondrial Dysfunction and Hypertension. Hypertension 2020, 76, 1980–1991. [Google Scholar] [CrossRef]
  420. Hool, L.C. The L-type Ca2+ channel as a potential mediator of pathology during alterations in cellular redox state. Heart Lung Circ. 2009, 18, 3–10. [Google Scholar] [CrossRef]
  421. Scragg, J.L.; Dallas, M.L.; Wilkinson, J.A.; Varadi, G.; Peers, C. Carbon monoxide inhibits L-type Ca2+ channels via redox modulation of key cysteine residues by mitochondrial reactive oxygen species. J. Biol. Chem. 2008, 283, 24412–24419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  422. Muralidharan, P.; Cserne Szappanos, H.; Ingley, E.; Hool, L. Evidence for redox sensing by a human cardiac calcium channel. Sci. Rep. 2016, 6, 19067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  423. Amberg, G.C.; Earley, S.; Glapa, S.A. Local regulation of arterial L-type calcium channels by reactive oxygen species. Circ. Res. 2010, 107, 1002–1010. [Google Scholar] [CrossRef] [PubMed]
  424. Chaplin, N.L.; Amberg, G.C. Hydrogen peroxide mediates oxidant-dependent stimulation of arterial smooth muscle L-type calcium channels. Am. J. Physiol. Cell. Physiol. 2012, 302, C1382–C1393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  425. Chaplin, N.L.; Nieves-Cintron, M.; Fresquez, A.M.; Navedo, M.F.; Amberg, G.C. Arterial Smooth Muscle Mitochondria Amplify Hydrogen Peroxide Microdomains Functionally Coupled to L-Type Calcium Channels. Circ. Res. 2015, 117, 1013–1023. [Google Scholar] [CrossRef] [PubMed]
  426. Ochi, R.; Dhagia, V.; Lakhkar, A.; Patel, D.; Wolin, M.S.; Gupte, S.A. Rotenone-stimulated superoxide release from mitochondrial complex I acutely augments L-type Ca2+ current in A7r5 aortic smooth muscle cells. Am. J. Physiol. Heart Circ. Physiol. 2016, 310, H1118–H1128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  427. Chiamvimonvat, N.; O’Rourke, B.; Kamp, T.J.; Kallen, R.G.; Hofmann, F.; Flockerzi, V.; Marban, E. Functional consequences of sulfhydryl modification in the pore-forming subunits of cardiovascular Ca2+ and Na+ channels. Circ. Res. 1995, 76, 325–334. [Google Scholar] [CrossRef]
  428. Fusi, F.; Saponara, S.; Gagov, H.; Sgaragli, G. 2,5-Di-t-butyl-1,4-benzohydroquinone (BHQ) inhibits vascular L-type Ca2+ channel via superoxide anion generation. Br. J. Pharmacol. 2001, 133, 988–996. [Google Scholar] [CrossRef] [Green Version]
  429. Sotnikova, R. Investigation of the mechanisms underlying H2O2-evoked contraction in the isolated rat aorta. Gen. Pharmacol. 1998, 31, 115–119. [Google Scholar] [CrossRef]
  430. Yang, Z.W.; Zheng, T.; Wang, J.; Zhang, A.; Altura, B.T.; Altura, B.M. Hydrogen peroxide induces contraction and raises [Ca2+]i in canine cerebral arterial smooth muscle: Participation of cellular signaling pathways. Naunyn-Schmiedeberg Arch. Pharmacol. 1999, 360, 646–653. [Google Scholar] [CrossRef]
  431. Tabet, F.; Savoia, C.; Schiffrin, E.L.; Touyz, R.M. Differential calcium regulation by hydrogen peroxide and superoxide in vascular smooth muscle cells from spontaneously hypertensive rats. J. Cardiovasc. Pharmacol. 2004, 44, 200–208. [Google Scholar] [CrossRef]
  432. Garcia-Redondo, A.B.; Briones, A.M.; Beltran, A.E.; Alonso, M.J.; Simonsen, U.; Salaices, M. Hypertension increases contractile responses to hydrogen peroxide in resistance arteries through increased thromboxane A2, Ca2+, and superoxide anion levels. J. Pharmacol. Exp. Ther. 2009, 328, 19–27. [Google Scholar] [CrossRef] [Green Version]
  433. Garcia-Redondo, A.B.; Briones, A.M.; Avendano, M.S.; Hernanz, R.; Alonso, M.J.; Salaices, M. Losartan and tempol treatments normalize the increased response to hydrogen peroxide in resistance arteries from hypertensive rats. J. Hypertens. 2009, 27, 1814–1822. [Google Scholar] [CrossRef]
  434. Santiago, E.; Contreras, C.; Garcia-Sacristan, A.; Sanchez, A.; Rivera, L.; Climent, B.; Prieto, D. Signaling pathways involved in the H2O2-induced vasoconstriction of rat coronary arteries. Free Radic. Biol. Med. 2013, 60, 136–146. [Google Scholar] [CrossRef]
  435. Garcia-Redondo, A.B.; Briones, A.M.; Martinez-Revelles, S.; Palao, T.; Vila, L.; Alonso, M.J.; Salaices, M. c-Src, ERK1/2 and Rho kinase mediate hydrogen peroxide-induced vascular contraction in hypertension: Role of TXA2, NAD(P)H oxidase and mitochondria. J. Hypertens. 2015, 33, 77–87. [Google Scholar] [CrossRef]
  436. Pan, B.X.; Zhao, G.L.; Huang, X.L.; Zhao, K.S. Mobilization of intracellular calcium by peroxynitrite in arteriolar smooth muscle cells from rats. Redox Rep. 2004, 9, 49–55. [Google Scholar] [CrossRef]
  437. Zimmerman, M.C.; Takapoo, M.; Jagadeesha, D.K.; Stanic, B.; Banfi, B.; Bhalla, R.C.; Miller, F.J., Jr. Activation of NADPH oxidase 1 increases intracellular calcium and migration of smooth muscle cells. Hypertension 2011, 58, 446–453. [Google Scholar] [CrossRef] [Green Version]
  438. Matoba, T.; Shimokawa, H.; Nakashima, M.; Hirakawa, Y.; Mukai, Y.; Hirano, K.; Kanaide, H.; Takeshita, A. Hydrogen peroxide is an endothelium-derived hyperpolarizing factor in mice. J. Clin. Investig. 2000, 106, 1521–1530. [Google Scholar] [CrossRef] [Green Version]
  439. Matoba, T.; Shimokawa, H.; Kubota, H.; Morikawa, K.; Fujiki, T.; Kunihiro, I.; Mukai, Y.; Hirakawa, Y.; Takeshita, A. Hydrogen peroxide is an endothelium-derived hyperpolarizing factor in human mesenteric arteries. Biochem. Biophys. Res. Commun. 2002, 290, 909–913. [Google Scholar] [CrossRef]
  440. Matoba, T.; Shimokawa, H.; Morikawa, K.; Kubota, H.; Kunihiro, I.; Urakami-Harasawa, L.; Mukai, Y.; Hirakawa, Y.; Akaike, T.; Takeshita, A. Electron spin resonance detection of hydrogen peroxide as an endothelium-derived hyperpolarizing factor in porcine coronary microvessels. Arterioscler. Thromb. Vasc. Biol. 2003, 23, 1224–1230. [Google Scholar] [CrossRef] [Green Version]
  441. Capettini, L.S.; Cortes, S.F.; Gomes, M.A.; Silva, G.A.; Pesquero, J.L.; Lopes, M.J.; Teixeira, M.M.; Lemos, V.S. Neuronal nitric oxide synthase-derived hydrogen peroxide is a major endothelium-dependent relaxing factor. Am. J. Physiol. Heart Circ. Physiol. 2008, 295, H2503–H2511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  442. Braunstein, T.H.; Inoue, R.; Cribbs, L.; Oike, M.; Ito, Y.; Holstein-Rathlou, N.H.; Jensen, L.J. The role of L- and T-type calcium channels in local and remote calcium responses in rat mesenteric terminal arterioles. J. Vasc. Res. 2009, 46, 138–151. [Google Scholar] [CrossRef] [PubMed]
  443. Shaifta, Y.; Snetkov, V.A.; Prieto-Lloret, J.; Knock, G.A.; Smirnov, S.V.; Aaronson, P.I.; Ward, J.P. Sphingosylphosphorylcholine potentiates vasoreactivity and voltage-gated Ca2+ entry via NOX1 and reactive oxygen species. Cardiovasc. Res. 2015, 106, 121–130. [Google Scholar] [CrossRef] [Green Version]
  444. Vogel, P.A.; Yang, X.; Moss, N.G.; Arendshorst, W.J. Superoxide enhances Ca2+ entry through L-type channels in the renal afferent arteriole. Hypertension 2015, 66, 374–381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  445. Fernando, V.; Zheng, X.; Walia, Y.; Sharma, V.; Letson, J.; Furuta, S. S-Nitrosylation: An Emerging Paradigm of Redox Signaling. Antioxidants 2019, 8, 404. [Google Scholar] [CrossRef]
  446. Hu, Z.; Zhang, B.; Lim, L.J.Y.; Loh, W.Z.K.; Yu, D.; Tan, B.W.Q.; Liang, M.C.; Huang, Z.; Leo, C.H.; Huang, H.; et al. S-Nitrosylation-Mediated Reduction of Cav1.2 Surface Expression and Open Probability Underlies Attenuated Vasoconstriction Induced by Nitric Oxide. Hypertension 2022, 79, 2854–2866. [Google Scholar] [CrossRef]
  447. Poteser, M.; Romanin, C.; Schreibmayer, W.; Mayer, B.; Groschner, K. S-nitrosation controls gating and conductance of the alpha 1 subunit of class C L-type Ca2+ channels. J. Biol. Chem. 2001, 276, 14797–14803. [Google Scholar] [CrossRef] [Green Version]
  448. Schulz, E.; Gori, T.; Munzel, T. Oxidative stress and endothelial dysfunction in hypertension. Hypertens. Res. 2011, 34, 665–673. [Google Scholar] [CrossRef]
  449. Kaludercic, N.; Deshwal, S.; Di Lisa, F. Reactive oxygen species and redox compartmentalization. Front. Physiol. 2014, 5, 285. [Google Scholar] [CrossRef] [Green Version]
  450. Gopalakrishna, R.; Anderson, W.B. Ca2+- and phospholipid-independent activation of protein kinase C by selective oxidative modification of the regulatory domain. Proc. Natl. Acad. Sci. USA 1989, 86, 6758–6762. [Google Scholar] [CrossRef] [Green Version]
  451. 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] [Green Version]
  452. Chakraborti, S.; Chakraborti, T. Down-regulation of protein kinase C attenuates the oxidant hydrogen peroxide-mediated activation of phospholipase A2 in pulmonary vascular smooth muscle cells. Cell. Signal. 1995, 7, 75–83. [Google Scholar] [CrossRef]
  453. Li, P.F.; Maasch, C.; Haller, H.; Dietz, R.; von Harsdorf, R. Requirement for protein kinase C in reactive oxygen species-induced apoptosis of vascular smooth muscle cells. Circulation 1999, 100, 967–973. [Google Scholar] [CrossRef]
  454. Steinberg, S.F. Mechanisms for redox-regulation of protein kinase C. Front. Pharmacol. 2015, 6, 128. [Google Scholar] [CrossRef] [Green Version]
  455. Sato, H.; Sato, M.; Kanai, H.; Uchiyama, T.; Iso, T.; Ohyama, Y.; Sakamoto, H.; Tamura, J.; Nagai, R.; Kurabayashi, M. Mitochondrial reactive oxygen species and c-Src play a critical role in hypoxic response in vascular smooth muscle cells. Cardiovasc. Res. 2005, 67, 714–722. [Google Scholar] [CrossRef]
  456. Thakali, K.; Davenport, L.; Fink, G.D.; Watts, S.W. Cyclooxygenase, p38 mitogen-activated protein kinase (MAPK), extracellular signal-regulated kinase MAPK, Rho kinase, and Src mediate hydrogen peroxide-induced contraction of rat thoracic aorta and vena cava. J. Pharmacol. Exp. Ther. 2007, 320, 236–243. [Google Scholar] [CrossRef] [Green Version]
  457. Park, H.J.; Shin, K.C.; Yoou, S.K.; Kang, M.; Kim, J.G.; Sung, D.J.; Yu, W.; Lee, Y.; Kim, S.H.; Bae, Y.M.; et al. Hydrogen peroxide constricts rat arteries by activating Na+-permeable and Ca2+-permeable cation channels. Free Radic. Res. 2019, 53, 94–103. [Google Scholar] [CrossRef]
  458. Zuhlke, R.D.; Pitt, G.S.; Deisseroth, K.; Tsien, R.W.; Reuter, H. Calmodulin supports both inactivation and facilitation of L-type calcium channels. Nature 1999, 399, 159–162. [Google Scholar] [CrossRef]
  459. Sun, Y.; Xu, J.; Minobe, E.; Shimoara, S.; Hao, L.; Kameyama, M. Regulation of the Cav1.2 cardiac channel by redox via modulation of CaM interaction with the channel. J. Pharmacol. Sci. 2015, 128, 137–143. [Google Scholar] [CrossRef] [Green Version]
  460. Dalton, T.P.; Shertzer, H.G.; Puga, A. Regulation of gene expression by reactive oxygen. Annu. Rev. Pharmacol. Toxicol. 1999, 39, 67–101. [Google Scholar] [CrossRef]
  461. Liu, H.; Colavitti, R.; Rovira, I.I.; Finkel, T. Redox-dependent transcriptional regulation. Circ. Res. 2005, 97, 967–974. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  462. Wang, W.Z.; Pang, L.; Palade, P. Angiotensin II causes endothelial-dependent increase in expression of Cav1.2 protein in cultured arteries. Eur. J. Pharmacol. 2008, 599, 117–120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  463. Wang, W.; Pang, L.; Palade, P. Angiotensin II upregulates Cav1.2 protein expression in cultured arteries via endothelial H2O2 production. J. Vasc. Res. 2011, 48, 67–78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  464. Tsai, C.T.; Wang, D.L.; Chen, W.P.; Hwang, J.J.; Hsieh, C.S.; Hsu, K.L.; Tseng, C.D.; Lai, L.P.; Tseng, Y.Z.; Chiang, F.T.; et al. Angiotensin II increases expression of alpha1C subunit of L-type calcium channel through a reactive oxygen species and cAMP response element-binding protein-dependent pathway in HL-1 myocytes. Circ. Res. 2007, 100, 1476–1485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  465. Cervenka, L.; Horacek, V.; Vaneckova, I.; Hubacek, J.A.; Oliverio, M.I.; Coffman, T.M.; Navar, L.G. Essential role of AT1A receptor in the development of 2K1C hypertension. Hypertension 2002, 40, 735–741. [Google Scholar] [CrossRef]
  466. Cervenka, L.; Vaneckova, I.; Huskova, Z.; Vanourkova, Z.; Erbanova, M.; Thumova, M.; Skaroupkova, P.; Opocensky, M.; Maly, J.; Chabova, V.C.; et al. Pivotal role of angiotensin II receptor subtype 1A in the development of two-kidney, one-clip hypertension: Study in angiotensin II receptor subtype 1A knockout mice. J. Hypertens. 2008, 26, 1379–1389. [Google Scholar] [CrossRef] [Green Version]
  467. Isidori, A.M.; Graziadio, C.; Paragliola, R.M.; Cozzolino, A.; Ambrogio, A.G.; Colao, A.; Corsello, S.M.; Pivonello, R.; Group, A.B.C.S. The hypertension of Cushing’s syndrome: Controversies in the pathophysiology and focus on cardiovascular complications. J. Hypertens. 2015, 33, 44–60. [Google Scholar] [CrossRef] [Green Version]
  468. Barbot, M.; Ceccato, F.; Scaroni, C. The Pathophysiology and Treatment of Hypertension in Patients With Cushing’s Syndrome. Front. Endocrinol. 2019, 10, 321. [Google Scholar] [CrossRef]
  469. Schewe, J.; Seidel, E.; Forslund, S.; Marko, L.; Peters, J.; Muller, D.N.; Fahlke, C.; Stolting, G.; Scholl, U. Elevated aldosterone and blood pressure in a mouse model of familial hyperaldosteronism with ClC-2 mutation. Nat. Commun. 2019, 10, 5155. [Google Scholar] [CrossRef] [Green Version]
  470. Narayanan, D.; Xi, Q.; Pfeffer, L.M.; Jaggar, J.H. Mitochondria control functional Cav1.2 expression in smooth muscle cells of cerebral arteries. Circ. Res. 2010, 107, 631–641. [Google Scholar] [CrossRef] [Green Version]
  471. Nowicki, P.T.; Flavahan, S.; Hassanain, H.; Mitra, S.; Holland, S.; Goldschmidt-Clermont, P.J.; Flavahan, N.A. Redox signaling of the arteriolar myogenic response. Circ. Res. 2001, 89, 114–116. [Google Scholar] [CrossRef] [Green Version]
  472. Cseko, C.; Bagi, Z.; Koller, A. Biphasic effect of hydrogen peroxide on skeletal muscle arteriolar tone via activation of endothelial and smooth muscle signaling pathways. J. Appl. Physiol. 2004, 97, 1130–1137. [Google Scholar] [CrossRef]
  473. Li, L.; Lai, E.Y.; Wellstein, A.; Welch, W.J.; Wilcox, C.S. Differential effects of superoxide and hydrogen peroxide on myogenic signaling, membrane potential, and contractions of mouse renal afferent arterioles. Am. J. Physiol. Renal. Physiol. 2016, 310, F1197–F1205. [Google Scholar] [CrossRef] [Green Version]
  474. Wagenfeld, L.; von Domarus, F.; Weiss, S.; Klemm, M.; Richard, G.; Zeitz, O. The effect of reactive oxygen species on the myogenic tone of rat ophthalmic arteries with and without endothelium. Graefe Arch. Clin. Exp. Ophthalmol. 2013, 251, 2339–2344. [Google Scholar] [CrossRef]
  475. Veerareddy, S.; Cooke, C.L.; Baker, P.N.; Davidge, S.T. Gender differences in myogenic tone in superoxide dismutase knockout mouse: Animal model of oxidative stress. Am. J. Physiol. Heart Circ. Physiol. 2004, 287, H40–H45. [Google Scholar] [CrossRef]
  476. Ungvari, Z.; Csiszar, A.; Huang, A.; Kaminski, P.M.; Wolin, M.S.; Koller, A. High pressure induces superoxide production in isolated arteries via protein kinase C-dependent activation of NAD(P)H oxidase. Circulation 2003, 108, 1253–1258. [Google Scholar] [CrossRef] [Green Version]
  477. Lai, E.Y.; Wellstein, A.; Welch, W.J.; Wilcox, C.S. Superoxide modulates myogenic contractions of mouse afferent arterioles. Hypertension 2011, 58, 650–656. [Google Scholar] [CrossRef] [Green Version]
  478. Lai, E.Y.; Solis, G.; Luo, Z.; Carlstrom, M.; Sandberg, K.; Holland, S.; Wellstein, A.; Welch, W.J.; Wilcox, C.S. p47(phox) is required for afferent arteriolar contractile responses to angiotensin II and perfusion pressure in mice. Hypertension 2012, 59, 415–420. [Google Scholar] [CrossRef] [Green Version]
  479. Kendrick, D.J.; Mishra, R.C.; John, C.M.; Zhu, H.L.; Braun, A.P. Effects of Pharmacological Inhibitors of NADPH Oxidase on Myogenic Contractility and Evoked Vasoactive Responses in Rat Resistance Arteries. Front. Physiol. 2021, 12, 752366. [Google Scholar] [CrossRef]
  480. Mironova, G.Y.; Mazumdar, N.; Hashad, A.M.; El-Lakany, M.A.; Welsh, D.G. Defining a role of NADPH oxidase in myogenic tone development. Microcirculation 2022, 29, e12756. [Google Scholar] [CrossRef]
  481. Gebremedhin, D.; Terashvili, M.; Wickramasekera, N.; Zhang, D.X.; Rau, N.; Miura, H.; Harder, D.R. Redox signaling via oxidative inactivation of PTEN modulates pressure-dependent myogenic tone in rat middle cerebral arteries. PLoS ONE 2013, 8, e68498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  482. 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] [PubMed] [Green Version]
  483. Phillips, S.A.; Sylvester, F.A.; Frisbee, J.C. Oxidant stress and constrictor reactivity impair cerebral artery dilation in obese Zucker rats. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2005, 288, R522–R530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  484. Butcher, J.T.; Goodwill, A.G.; Stanley, S.C.; Frisbee, J.C. Differential impact of dilator stimuli on increased myogenic activation of cerebral and skeletal muscle resistance arterioles in obese zucker rats. Microcirculation 2013, 20, 579–589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  485. Velmurugan, G.V.; Sundaresan, N.R.; Gupta, M.P.; White, C. Defective Nrf2-dependent redox signalling contributes to microvascular dysfunction in type 2 diabetes. Cardiovasc. Res. 2013, 100, 143–150. [Google Scholar] [CrossRef]
  486. Springo, Z.; Tarantini, S.; Toth, P.; Tucsek, Z.; Koller, A.; Sonntag, W.E.; Csiszar, A.; Ungvari, Z. Aging Exacerbates Pressure-Induced Mitochondrial Oxidative Stress in Mouse Cerebral Arteries. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 2015, 70, 1355–1359. [Google Scholar] [CrossRef] [Green Version]
  487. Spurrell, B.E.; Murphy, T.V.; Hill, M.A. Tyrosine phosphorylation modulates arteriolar tone but is not fundamental to myogenic response. Am. J. Physiol. Heart Circ. Physiol. 2000, 278, H373–H382. [Google Scholar] [CrossRef] [Green Version]
  488. Murphy, T.V.; Spurrell, B.E.; Hill, M.A. Tyrosine phosphorylation following alterations in arteriolar intraluminal pressure and wall tension. Am. J. Physiol. Heart Circ. Physiol. 2001, 281, H1047–H1056. [Google Scholar] [CrossRef]
  489. Xiao, D.; Buchholz, J.N.; Zhang, L. Pregnancy attenuates uterine artery pressure-dependent vascular tone: Role of PKC/ERK pathway. Am. J. Physiol. Heart Circ. Physiol. 2006, 290, H2337–H2343. [Google Scholar] [CrossRef]
  490. Osol, G.; Laher, I.; Cipolla, M. Protein kinase C modulates basal myogenic tone in resistance arteries from the cerebral circulation. Circ. Res. 1991, 68, 359–367. [Google Scholar] [CrossRef] [Green Version]
  491. Hill, M.A.; Falcone, J.C.; Meininger, G.A. Evidence for protein kinase C involvement in arteriolar myogenic reactivity. Am. J. Physiol. 1990, 259, H1586–H1594. [Google Scholar] [CrossRef]
  492. Kizub, I.V.; Pavlova, O.O.; Johnson, C.D.; Soloviev, A.I.; Zholos, A.V. Rho kinase and protein kinase C involvement in vascular smooth muscle myofilament calcium sensitization in arteries from diabetic rats. Br. J. Pharmacol. 2010, 159, 1724–1731. [Google Scholar] [CrossRef] [Green Version]
  493. Novokhatska, T.; Tishkin, S.; Dosenko, V.; Boldyriev, A.; Ivanova, I.; Strielkov, I.; Soloviev, A. Correction of vascular hypercontractility in spontaneously hypertensive rats using shRNAs-induced delta protein kinase C gene silencing. Eur. J. Pharmacol. 2013, 718, 401–407. [Google Scholar] [CrossRef]
  494. Csato, V.; Peto, A.; Koller, A.; Edes, I.; Toth, A.; Papp, Z. Hydrogen peroxide elicits constriction of skeletal muscle arterioles by activating the arachidonic acid pathway. PLoS ONE 2014, 9, e103858. [Google Scholar] [CrossRef] [Green Version]
  495. Shaikh, A. A Practical Approach to Hypertension Management in Diabetes. Diabetes Ther. 2017, 8, 981–989. [Google Scholar] [CrossRef]
  496. Tocci, G.; Desideri, G.; Roca, E.; Calcullo, C.; Crippa, M.; De Luca, N.; Gaudio, G.V.; Lonati, L.M.; Orselli, L.; Scuteri, A.; et al. How to Improve Effectiveness and Adherence to Antihypertensive Drug Therapy: Central Role of Dihydropyridinic Calcium Channel Blockers in Hypertension. High Blood Press. Cardiovasc. Prev. 2018, 25, 25–34. [Google Scholar] [CrossRef] [Green Version]
  497. Odigboegwu, O.; Pan, L.J.; Chatterjee, P. Use of Antihypertensive Drugs During Preeclampsia. Front. Cardiovasc. Med. 2018, 5, 50. [Google Scholar] [CrossRef] [Green Version]
  498. Ojha, U.; Ruddaraju, S.; Sabapathy, N.; Ravindran, V.; Worapongsatitaya, P.; Haq, J.; Mohammed, R.; Patel, V. Current and Emerging Classes of Pharmacological Agents for the Management of Hypertension. Am. J. Cardiovasc. Drugs 2022, 22, 271–285. [Google Scholar] [CrossRef]
  499. Forman, H.J.; Zhang, H. Targeting oxidative stress in disease: Promise and limitations of antioxidant therapy. Nat. Rev. Drugs Discov. 2021, 20, 689–709. [Google Scholar] [CrossRef]
Figure 1. Topology of Cav1.2. Cav1.2 is formed by the pore-forming transmembrane α1c subunit, the intracellular β-subunit, the extracellular α2 subunit linked to the transmembrane δ1 subunit in smooth muscle cells.
Figure 1. Topology of Cav1.2. Cav1.2 is formed by the pore-forming transmembrane α1c subunit, the intracellular β-subunit, the extracellular α2 subunit linked to the transmembrane δ1 subunit in smooth muscle cells.
Antioxidants 11 02432 g001
Figure 2. Proposed mechanisms for regulating myogenic tone in small arteries and arterioles. Detailed information is discussed in the text. GPCR, G protein-coupled receptor, TRPV4, Transient receptor potential vanilloid-type 4; BKCa, large-conductance Ca2+-activated K+ channel; SR, sarcoplasmic reticulum; IP3, inositol triphosphate; RyR, ryanodine receptor; CaM, calmodulin; MLCK, myosin light-chain kinase; MLC20, 20 kD myosin light-chain.
Figure 2. Proposed mechanisms for regulating myogenic tone in small arteries and arterioles. Detailed information is discussed in the text. GPCR, G protein-coupled receptor, TRPV4, Transient receptor potential vanilloid-type 4; BKCa, large-conductance Ca2+-activated K+ channel; SR, sarcoplasmic reticulum; IP3, inositol triphosphate; RyR, ryanodine receptor; CaM, calmodulin; MLCK, myosin light-chain kinase; MLC20, 20 kD myosin light-chain.
Antioxidants 11 02432 g002
Figure 3. ROS generation by NOXs and mitochondria and detoxification. In vascular smooth muscle cells, ROS is primarily produced by NOXs and mitochondria. NOXs catalyze the production of O2•− by transferring one electron to O2 from NADPH. In mitochondria, O2•− is also produced from leaked electrons by Complexes I and III in the electronic transfer chain (ETC) during the electronic transfer. O2•− is dismutated to H2O2 by superoxide dismutase (SOD). H2O2 is subsequently decomposed to H2O by catalase, glutathione peroxidase (GPX) and peroxiredoxin (PRX). O2•− can be released into the cytosol from mitochondria via the voltage-dependent anion channel (VADC) on the mitochondrial outer membrane, whereas H2O2 in mitochondria can be transported to the cytosol via diffusion or via aquaporins. O2•− reacts with nitro oxide (NO) to produce peroxynitrite (ONOO), whereas H2O2 reacts with Fe2+ via the Fenton reaction to yield OH.
Figure 3. ROS generation by NOXs and mitochondria and detoxification. In vascular smooth muscle cells, ROS is primarily produced by NOXs and mitochondria. NOXs catalyze the production of O2•− by transferring one electron to O2 from NADPH. In mitochondria, O2•− is also produced from leaked electrons by Complexes I and III in the electronic transfer chain (ETC) during the electronic transfer. O2•− is dismutated to H2O2 by superoxide dismutase (SOD). H2O2 is subsequently decomposed to H2O by catalase, glutathione peroxidase (GPX) and peroxiredoxin (PRX). O2•− can be released into the cytosol from mitochondria via the voltage-dependent anion channel (VADC) on the mitochondrial outer membrane, whereas H2O2 in mitochondria can be transported to the cytosol via diffusion or via aquaporins. O2•− reacts with nitro oxide (NO) to produce peroxynitrite (ONOO), whereas H2O2 reacts with Fe2+ via the Fenton reaction to yield OH.
Antioxidants 11 02432 g003
Figure 4. Regulation of Cav1.2 by ROS microdomains. In vascular smooth muscle cells, ROS-derived from NOXs in plasma membrane and/or subcellular mitochondria in the immediate vicinity of Cav1.2 channels form ROS microdomains. Cav1.2 activity can be altered directly and indirectly by ROS. Within ROS microdomains, ROS can directly target cysteine residues in Cav1.2 to modify channel activity. In addition, ROS can also trigger ROS-dependent activation of PKC and/or c-Src. Activated PKC and c-Src in turn phosphorylate Cav1.2 and enhance channel activity.
Figure 4. Regulation of Cav1.2 by ROS microdomains. In vascular smooth muscle cells, ROS-derived from NOXs in plasma membrane and/or subcellular mitochondria in the immediate vicinity of Cav1.2 channels form ROS microdomains. Cav1.2 activity can be altered directly and indirectly by ROS. Within ROS microdomains, ROS can directly target cysteine residues in Cav1.2 to modify channel activity. In addition, ROS can also trigger ROS-dependent activation of PKC and/or c-Src. Activated PKC and c-Src in turn phosphorylate Cav1.2 and enhance channel activity.
Antioxidants 11 02432 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hu, X.-Q.; Zhang, L. Oxidative Regulation of Vascular Cav1.2 Channels Triggers Vascular Dysfunction in Hypertension-Related Disorders. Antioxidants 2022, 11, 2432. https://doi.org/10.3390/antiox11122432

AMA Style

Hu X-Q, Zhang L. Oxidative Regulation of Vascular Cav1.2 Channels Triggers Vascular Dysfunction in Hypertension-Related Disorders. Antioxidants. 2022; 11(12):2432. https://doi.org/10.3390/antiox11122432

Chicago/Turabian Style

Hu, Xiang-Qun, and Lubo Zhang. 2022. "Oxidative Regulation of Vascular Cav1.2 Channels Triggers Vascular Dysfunction in Hypertension-Related Disorders" Antioxidants 11, no. 12: 2432. https://doi.org/10.3390/antiox11122432

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