• Open Access
    Review

    Reactive oxygen species in cardiovascular diseases: an update

    Juanjuan Fei 1
    Laurie J. Demillard 2
    Jun Ren 1,3*

    Explor Med. 2022;3:188–204 DOI: https://doi.org/10.37349/emed.2022.00085

    Received: January 06, 2022 Accepted: February 27, 2022 Published: April 26, 2022

    Academic Editor: Milan Obradovic, University of Belgrade, Serbia

    This article belongs to the special issue Reactive Oxygen Species (ROS) in Pathophysiological Conditions

    Abstract

    Cardiovascular diseases are among the leading causes of death worldwide, imposing major health threats. Reactive oxygen species (ROS) are one of the most important products from the process of redox reactions. In the onset and progression of cardiovascular diseases, ROS are believed to heavily influence homeostasis of lipids, proteins, DNA, mitochondria, and energy metabolism. As ROS production increases, the heart is damaged, leading to further production of ROS. The vicious cycle continues on as additional ROS are generated. For example, recent evidence indicated that connexin 43 (Cx43) deficiency and pyruvate kinase M2 (PKM2) activation led to a loss of protection in cardiomyocytes. In this context, a better understanding of the mechanisms behind ROS production is vital in determining effective treatment and management strategies for cardiovascular diseases.

    Keywords

    Reactive oxygen species, cardiovascular diseases, nicotinamide adenine dinucleotide phosphate oxidase, xanthine oxidase, monoamine oxidases, nitric oxide synthase

    Introduction

    Reactive oxygen species (ROS) play an important role in the pathophysiology of cardiovascular dysfunction [13]. An increased production of ROS is associated with the development of an imbalance of generation and elimination in redox reactions [4]. The heart produces, nicotinamide adenine dinucleotide phosphate (NADPH) oxidase, xanthine oxidase (XO), uncoupled nitric oxide (NO) synthase (NOS), and monoamine oxidases (MAOs) by oxidant systems [5]. Oxidative stress performs different functions based on the amount produced. Under normal conditions, oxidative stress is maintained at low levels to sustain physiological metabolism [3, 5, 6]. However, excessive oxidative stress can damage the cardiovascular system. ROS have various effects on cardiovascular diseases that lead to cell regeneration defect, lipid peroxidation, protein degeneration, DNA damage, mitochondrial injury and energy metabolism disorder [5, 711].

    ROS discovery and components

    After the discovery of free radicals in biological systems in 1950 [12], free radicals have been reported to be involved in diverse pathological processes. Redox signaling is an essential process of an organism. ROS include oxygen radicals, such as superoxide anion (O2), hydroxyl radical (OH), hydrogen peroxide (H2O2), NO, peroxynitrite (ONOO) and hypochlorite (OCL), which are highly reactive molecules [13, 14].

    Biological function of ROS

    In nature, many oxidase enzymes contribute to ROS generation. Oxidative stress is characterized by an imbalance between the pro-oxidant and antioxidant systems, resulting in an increase of ROS production [11, 15] (Figure 1). In vivo, there are many oxidant systems such as NADPH oxidases (NOXs), mitochondrial respiratory chain enzymes, XO, MAOs, uncoupled endothelial NOS (eNOS), and lipoxygenases. These systems influence the formation and development of cardiovascular diseases through ROS production [1621]. All these oxidases can be regulated by antioxidant systems, including superoxide dismutase (SOD), catalase, glutathione (GSH) peroxidases, paraoxonases, thioredoxin system, and peroxiredoxins [21]. In addition, mitochondrial mutations can lead to ROS production [22]. Mitochondrial DNA (mtDNA) is easily damaged because of its limited capacity to repair DNA [22]. In addition, excessive mitochondrial ROS (mtROS) generation results in increased possibility of permeability transition pore (PTP) opening and leads to cell death [23]. The increased ROS also leads to mtDNA leakage and contributes to inflammation [24]. Furthermore, ROS contributes to dysregulation of intracellular Ca2+ homeostasis, by causing mitochondrial membrane depolarization and Ca2+ release [25].

    Multiple oxidant systems contribute to the production of ROS

    Oxidant systems

    NOX

    NOXs are multi-transmembrane enzyme complexes composed of a plasma membrane and cytosolic components [26, 27]. They are also the major enzymes involved in the generation of ROS that contribute to cardiovascular diseases [28] (Figure 1). NOXs consist of seven isoforms NOXs: NOX1, NOX2/gp91(phox), NOX3, NOX4, NOX5, dual oxidase 1 (DUOX1) and dual oxidase 2 (DUOX2) [26, 29]. Each oxidase expresses itself differently in cardiovascular, endothelial, and vascular smooth muscle cells [3032]. Activated NOXs can utilize NADPH as an electron donor and transfer an electron to molecular oxygen. This reaction generates a superoxide molecule that plays an important role in the redox reaction [23, 27, 33, 34]. Several studies have identified that NOXs contribute to cardiovascular diseases, such as atherosclerosis, hypertension, heart failure, and ischemia-reperfusion injury (I/R) [30]. The increased expression and activation of NOX can contribute to ROS production [1, 3537]. ROS signaling is vital in establishing communication between mitochondria and NOXs in the cardiovascular system [3, 13]. Further research indicates that NOX can also activate the nucleotide-binding and oligomerization domain (NOD)-like receptor family pyrin domain containing 3 (NLRP3) inflammasome in macrophages which contributes to the progression of atherosclerosis [38, 39].

    Mitochondrial respiratory chain

    The structure of the mitochondrion was first described in 1888. Mitochondria are known as the source of chemical energy within a cell, but also participate in aerobic respiration which generates ROS [40, 41]. mtROS production has been observed in vivo at complex I and complex III in the electron transport chain (ETC) [4244]. The process of electron transfer generates mtROS that play an important role in the intracellular redox state [38, 45]. Under physiological conditions, the loss of electrons is minimal. However, during conditions of oxidative stress, the production of mtROS is increased and causes damage to mitochondria [46]. This reaction indicates that mitochondria themselves may also be susceptible to the overexpression of ROS they produce. These effects cause damage to mitochondrial DNA and membranes, further impairing the normal activity of the ETC and generating ROS through positive feedback [47]. In addition, the changes in mtROS impact K+ and Ca2+ channels that influence cell functions [48, 49].

    XO

    XO is a cytoplasmic enzyme that can be converted from xanthine dehydrogenase [38]. XO catalyzes the oxidation of hypoxanthine to xanthine by the transfer of an electron to oxygen and produces superoxide [23, 30] (Figure 1). Allopurinol is an inhibitor of XO that can be used to reduce levels of uric acid [50]. Recently, allopurinol has been found to have potential cardiovascular protection by modulating ROS and Ca2+ [50, 51].

    MAOs

    MAOs are located at the outer membrane of mitochondria [52]. There are two isoforms of MAO namely MAO-A and MAO-B [23]. Under pathological conditions, the increased activation of MAOs generates excessive H2O2 and aldehyde, leading to dysfunction of mitochondrion [53, 54] (Figure 1). MAO is also an oxidase which has been closely associated with cardiovascular diseases such as vascular dysfunction, I/R, maladaptive hypertrophy and heart failure [5557]. Some research indicates that the increased activation of MAOs disturbs the balance of redox and leads to excessive production of ROS that may damage cardiomyocytes [23]. Moreover, MAOs can increase the production of mtROS resulting in the activation of inflammasome [23].

    Uncoupled eNOS

    NOS consists of three isoforms: eNOS, neuronal NOS (nNOS) and inducible NOS (iNOS) [30] (Figure 1). Under physiological conditions, eNOS maintains the balance of endothelial function by binding to L-arginine with the assistance of tetrahydrobiopterin (BH4) during NO synthesis [58]. BH4 is an important element for maintaining the stability of eNOS and is also the basic cofactor for NO synthesis [58]. However, during oxidative stress, BH4 will be converted to dihydrobiopterin (BH2) and promotes uncoupling by interacting with eNOS [59, 60].

    The role of ROS in cardiovascular diseases

    ROS play an important role in the pathogenesis of cardiovascular diseases, such as I/R, vascular endothelial and atherosclerosis, hypertension, diabetic cardiomyopathy (DCM), heart failure, cardiac arrhythmias, and aortic aneurysms [3, 42].

    I/R

    A large number of ROS are produced during the process of I/R, due to activation of the ETC and several enzymes [54]. ROS also accelerate the loss of adenosine triphosphate (ATP) during the period of I/R [61]. In early reperfusion, the production of ROS exceeds the removal capacity of antioxidant systems leading to mitochondrial respiratory complex peroxidation [30] (Figure 2). These changes lead to oxidative damage and cardiomyocyte death [6, 61, 62].

    The development of cardiovascular diseases is caused by an imbalance of redox system. GPx: GSH peroxidase; Trx: thioredoxin

    During the period of hypoxia/reoxygenation, the changes of intracellular calcium in aortic endothelial cells impact the uptake of Ca2+ in mitochondria [22] (Figure 2). The generation of Ca2+ can be modulated by ROS derived from the NOX [22]. Research indicates that the upregulation of receptor-interacting protein 3 (Ripk3) increases production of mtROS and cell death in I/R [63]. The increased production of mtROS is modulated by Ca2+ overload and XO in the overexpression of Ripk3 [63, 64] (Table 1, Figure 3). In combination with oxidative stress, calcium surplus leads to the opening of mitochondrial permeability transition pore (mPTP) and cardiomyocyte death [6, 22]. In addition to the Ca2+ overload, mitochondrial dysfunction and ROS-induced ROS release (RIRR) also promote cell death [61]. The release of ROS can be reduced by activation of adenosine monophosphate-activated protein kinase (AMPK)/Akt/glycogen synthase kinase-3β (GSK-3β) pathway in I/R [65] (Table 1). Inhibition of transforming growth factor-activated kinase 1 (TAK1) can also reduce ROS production in I/R [66] (Table 1). Moreover, evidence shows that signal transducer and activator of transcription 3 (STAT3) is able to control ROS production [67]. During ischemia, STAT3 overexpression reduces the production of superoxide in the heart [61, 68]. Activation of STAT3 can improve resistance in I/R and prevent cardiac remodeling by modulating interleukin-11 (IL-11) [69] (Figure 3). Additionally, nicotinamide phosphoribosyltransferase (Nampt) was proven to be effective for protection in I/R [36, 70]. Nampt inhibits apoptosis by regulating the level of nicotinamide adenine dinucleotide (NAD+) and ATP [71]. Additionally, mitochondrial connexin 43 (Cx43) modulates the production of mtROS [6, 72]. Cx43 protects the function of cardiomyocytes by casein kinase 1 (CK1) which leads to the phosphorylation of Cx43 [73]. Recently, research findings demonstrated that Cx43 deficiency can lead to a loss of protection in I/R [6, 72, 74, 75] (Figure 3). Furthermore, the citric acid cycle (CAC) is an important element of metabolism. However, it has been reported to be associated with mtROS production recently [76]. The research indicates that succinate was increased in ischemia and then oxidized by succinate dehydrogenase during the period of reperfusion leading to mtROS production [76] (Figure 3). Iron-catalysed reactions are another condition which causes increased ROS production [77]. However, the role of Fe–S clusters is an area for further research [49]. Recent evidence indicates that at the end of I/R, activation of hypoxia-inducible transcription factor (HIF) can increase levels of mitochondrial NADPH and reduce levels of mtROS which prevent cardiac fibroblast formation [78, 79] (Figure 3). Similarly, Gap junction protein Alpha 1-20 kDa (GJA1-20k) and pleckstrin homology-like domain, family A, member 1 (PHLDA1) are also two important factors capable of reducing ROS [80, 81] (Table 1, Figure 3). GPx is also found to inhibit left ventricular (LV) remodeling and improve cardiac tissue survival in I/R [36] (Figure 2, Figure 3). Glutathione exhibits protective effects in cardiomyocytes by reducing the formation of ROS [82] (Figure 3). The reduction of ROS is essential in providing protection during I/R.

    Response of ROS under different mechanisms in cardiovascular diseases

    Disease typesROSMechanismsPMIDsPublication dates
    I/RRipk3–Ca2+ overload–XO–ROS295020452018-02
    HIF-1–NADPH–mtROS347638602022-02
    AMPK–Akt–GSK-3β–ROS281283612017-01
    TAK1↓–ROS323782872020-10
    GJA1-20k–ROS346088632021-10
    PHLDA1–ROS319816282020-03
    HypertensionJMJD1A–ROS324619962020-05
    AMPK–PAR1–ROS292877252018-01
    AMPK–PINK1–parkin–ROS292856902018-03
    AMPK–O-GlcNAC↓–ROS292856902018-01
    Sirt1–LKB1–MAPK–ROS237075582013-10
    Foxo1–SOD2–ROS306775122019-06
    ROS–Nrf2–ARE–ROS336569042020-12
    AngII–ROS306439682019-01
    CELF1↓–PEBP1–MAPK↓–ROS346690212022-01
    AtherosclerosisIRS-1–ROS330002672020-11
    HIF-1–ROS351110452021-12
    PON2–ROS174041542007-04
    AMPK–NOX↓–ROS313311112019-07
    PKM2↓–G6P–NADPH–ROS302221362018-10
    Sirt3–Foxo3α–MnSOD–ROS236653962013-10
    Sirt3–IDH2–GSH–ROS304553812019-01
    DCMRAGE–NOX–ROS279166502017-04
    PKC–NF-κB–iNOS–ROS279166502017-04
    NEU1–AMPKα↓–Sirt3↓–SOD2↓–ROS350025282022-01
    Heart failureSirt3–CypD–mPTP–SOD↓–ROS335084342021-03
    Vascular endothelialSirt2–Foxo3α–SOD–ROS340281772021-07
    Display full size

    AngII: angiotensin II; ARE: antioxidant response element; CELF1: cytimidine uracil guanine triplet repeat-binding protein 1; CypD: cyclophilin D; Foxo1: forkhead box protein O1; Foxo3α: forkhead box transcription factor 3α; G6P: glucose-6-phosphate; IDH2: isocitrate dehydrogenase 2; IRS-1: insulin receptor substrate 1; JMJD1A: Jumonji domain containing 1A; LKB1: liver kinase B1; MAPK: mitogen-activated protein kinases; MnSOD: manganese SOD; NEU1: neuraminidase 1; NF-κB: nuclear factor kappaB; Nrf2: nuclear factor E2-related factor 2; O-GlcNAC: O-linked N-acetylglucosamine; PAR1: protease-activated receptor 1; PEBP1: phosphatidylethanolamine binding protein 1; PINK1: phosphatase and tensin homolog-induced putative kinase 1; PKC: protein kinase C; PKM2: pyruvate kinase M2; PON2: paraoxonase-2; RAGE: receptor for advanced glycation end products; Ripk3: receptor-interacting serine/threonine-protein kinase 3; Sirt 1: sirtuin 1; PMID: PubMed ID; ↑: increase; ↓: decrease

    The mechanisms of ROS in cardiovascular diseases. NAC: N-acetylcysteine; OPLAH: oxoprolinase; Vc: vitamin C; ↑: increase; ↓: decrease

    Vascular endothelial and atherosclerosis

    The process of atherosclerosis is accelerated by various factors, such as the generation of ROS, inflammatory signaling, and endothelium dysfunction [21]. However, atherosclerosis begins with endothelium dysfunction and plays an important role in vascular homeostasis [21] (Figure 2). The intima of the endothelium is formed of one layer of endothelial cells surrounded by adhesion molecules [83]. In physiological conditions, steady blood flow has little effect on vascular endothelium and increases the production of NO [21]. When the endothelium is injured, lipids deposit in the vascular wall. This causes a change in vascular blood flow and upregulates NOX, leading to oxidative stress [21]. The abnormal production of NO is caused by eNOS disorder [84]. Meanwhile, the reduction and low bioavailability of NO contributes to atherosclerosis because N-hexanoyl-D-erythro-sphingosine is activated [85]. Moreover, Sirt2 plays an important role in maintaining the function of endothelial cells. It can reduce ROS production via Sirt2/Foxo3α/SOD pathway [86] (Table 1). Therefore, the improvement of endothelial cell function could have a protective effect on vascular function [87]. Sirt3 is also the member of sirtuin family which has a protective effect in cardiovascular diseases (Figure 3). It can mediate Foxo3α/MnSOD and IDH2/GSH pathways to reduce ROS production [8890]. HIF-1 is a heterodimeric protein which plays an important role in atherosclerosis by regulating ROS and NO production [91] (Figure 3). Moreover, novel researches demonstrated that PKM2 is related to release of ROS. The inhibition of PKM2 can activate the G6P/NADPH pathway to reduce ROS in cardiomyocytes exposed to oxygen/glucose [92, 93] (Figure 3). Apelin/APJ is a member of G protein-coupled receptors (GPCR) and is expressed on endothelial and smooth muscle cells [94]. Apelin-13 reduces lipid accumulation of foam cells through activating class III phosphatidylinositol 3-kinase (PI3K)/Beclin-1 pathway [9496] (Figure 3). Kruppel-like factor 2 (KLF2) and KLF4 have also been reported to act as protective factors in atherosclerosis [21, 97]. Deficiencies of KLF2 and KLF4 accelerate the atherosclerotic process by inducing eNOS [98]. NF-κB can also decrease ROS accumulation by downregulation of c-Jun N-terminal kinase (JNK) [99] (Figure 3). Research has shown that loss of insulin signaling (IRS-1) in vascular endothelium leads to endothelial dysfunction and atherosclerosis [100, 101] (Figure 3). The research found that overexpression of PON2 would reduce the production of ROS by decreasing endoplasmic reticulum stress [102] (Figure 3). Furthermore, the downregulation of receptor-interacting serine/Ripk3 can reduce the activation of inflammatory processes, which has a protective effect in atherosclerosis [103]. The factors HIF-1 and AMPK also participate in ROS production in the process of atherosclerosis [91, 104, 105].

    Hypertension

    There is robust evidence that ROS production is increased in patients with hypertension [14]. Mechanisms contributing to hypertension are complex; oxidative stress is one of the most important factors [106]. Uncoupled eNOS can promote an increase in cardiomyocyte hypertrophy in response to chronic hypertension [107, 108]. Cardiomyocyte hypertrophy increases the expression of XO [27] (Figure 2). CD8+ T cells contribute to salt-sensitive hypertension by increasing ROS production and sodium retention [109]. Additionally, a reduction in Sirt3 leads to endothelial dysfunction by creating mitochondrial oxidative stress in hypertension [14] (Figure 3). The depletion of Sirt3 contributes to vascular inflammation and hypertrophy, leading to progression of hypertension [110]. The activation via Sirt1/LKB1 and CELF1/PEBP1 pathways can both reduce ROS production [111, 112]. AMPK is a key factor in reducing the production of ROS via different pathways such as AMPK/PAR1, AMPK/phosphatase and tensin homolog-induced putative kinase 1 (PINK1)/parkin and AMPK/O-GlcNAC [113116] (Table 1). ROS can also be reduced through activation of JMJD1A and Foxo1 [117, 118] (Table 1). Furthermore, NOX4-mediated mtROS signaling is important in the response to chronic pressure overload [27]. NOX4 protects the heart from hypertrophic dysfunction by activation of HIF1a/vascular endothelial growth factor (VEGF) signaling pathway in cardiomyocytes [119]. In hypertension models, the activation of Nrf2 demonstrated an antihypertensive effect [120] (Figure 3). The activation of the Nrf2 pathway induces the production of antioxidant enzymes [120] (Table 1). Nrf2 can be modulated by the mammalian STE20-like protein kinases 1/2 (Mst1/2) to sustain the balance of cellular redox reactions [38, 121]. Thus, presence of mitochondrial antioxidants may improve vascular function [14]. Finally, the Trx system is able to reduce ROS and act as an anti-hypertrophic factor [27] (Figure 3).

    DCM

    Hyperglycemia causes an alteration of mitochondrial morphology, including mitochondrial fragmentation and swelling, leading to an increase in ROS production [60, 122, 123]. Increased production of ROS can also contribute to changes in mitochondrial morphology [23]. However, antioxidants may reduce the production of ROS, offering a potential therapeutic strategy for the treatment of DCM [124]. Altered mitochondrial function may inhibit insulin signaling which leads to activation of PKC [125]. PKC is from a family of kinases related to an increase in ROS production [101]. PKC activates NOX in diabetes and induces oxidative stress [59] (Figure 3). PKC can also activate NF-κB to increase ROS release [59] (Table 1). Furthermore, PKC can be up-regulated by p66shc, and then inhibit eNOS activity, creating a vicious cycle [101] (Figure 2). AMPK is also reported to reduce the production of mtROS [126] (Figure 3). Activated phosphorylated AMPK (pAMPK) can inhibit pyroptosis in DCM [127]. However, research indicates that inhibiting the expression of AMPK/p38 MAPK signaling reduces ROS production [128]. This production of ROS is different due to the distinct pathway that AMPK activates. For example, activation of NEU1 inhibits the AMPKα/Sirt3-SOD2 pathway and leads to ROS production [129] (Table 1). Moreover, advanced glycation end products (AGE) binding to the RAGE cause activation of NADPH oxidase enzymes that lead to ROS generation [59] (Table 1, Figure 3). In DCM, the increased expression of NOX2 contributes to ROS production [23]. ROS formation through NOX is associated with pathways involving sodium/glucose cotransporter 1 (SGLT1), PKCβ, and calcium/calmodulin dependent kinase II (CaMKII) [130]. Further, the decreased activation of NOX2 can reduce myocardial oxidative stress and remodeling which improves cardiac function [131]. In addition to ROS, high glucose-induced generation of NOS causes DNA damage [23, 60]. Excess glucose also induces arginase activity and upregulates eNOS activity [23, 60]. The expression of NOS is increased in diabetic hearts and leads to enhanced lipid peroxidation and peroxynitrite generation [23]. Lastly, iNOS is up-regulated in DCM, with increased levels of 4-hydroxynonenal (4-HNE) [59]. Supplementation with BH4 may be possible to reduce oxidative stress in DCM [59] (Figure 3). Coenzyme-Q10 (CQ10) mitochondria-targeted antioxidants also reduce H2O2 in hyperglycemia [59] (Figure 3).

    Heart failure

    Heart failure is a condition where the heart exhibits abnormal cardiac structure and function leading to pump failure [22, 27]. The mechanisms contributing to heart failure are complex, including mitochondrial dysfunction, redox imbalance, ion disorder, and inflammation [24]. The expressions of XO and MAO are elevated in heart failure, leading to increased ROS production [27, 36] (Figure 2). One of the major mechanisms in heart failure is an increase of mtROS due to mitochondrial stress [24] (Figure 3). Furthermore, ROS release contributes to the progression of heart failure and leads to cardiac dysfunction and ventricular remodeling [132]. In heart failure, diastolic calcium leak contributes to cytoplasmic calcium overload and diastolic dysfunction and arrhythmia [27]. In turn, ROS inhibits calcium reuptake and impacts diastolic function [133]. Additionally, ROS activates the apoptosis signal-regulating kinase-1 (ASK-1)/JNK-dependent pathway which causes apoptosis in in vivo models of heart failure [134]. Sirt3 has been shown to be downregulated in the failing heart [24] (Figure 2). Sirt3 offers heart protection by maintaining mitochondria function [135]. The activity of Sirt3 is mediated by NAD+ availability [136]. Decreased NAD+ levels suppress NAD+ dependent protein deacetylation, resulting in mitochondrial protein hyperacetylation and impaired function [24, 136]. Sirt3 can also regulate ROS production by CypD/mPTP/SOD pathway [137] (Table 1).

    Several studies demonstrated that antioxidant NAC can improve GSH levels, reduce ROS production, and improve cardiac function [36, 59] (Figure 3). A number of clinical trials have demonstrated that the enhanced expression of 5-oxoprolinase (OPLAH) could improve GSH/oxidized glutathione (GSSG) ratio and benefit heart failure [36] (Figure 3). Vc may also be used as an antioxidant to improve endothelial function in heart failure [22] (Figure 3).

    Cardiac arrhythmias

    Atrial fibrillation (AF) is one of the common arrhythmias related to ATP deficiency and changes in Na+, K+, and Ca2+ channels [138] (Figure 3). Recently ROS was reported to play an important role in AF [139]. NOX2-derived ROS generation has been implicated in experimental and clinical AF [27]. The production of mtROS contributes to cardiac fibrosis which is a characteristic of AF [139]. Furthermore, AF causes the opening of mPTP leading to a disruption of Ca2+ homeostasis and mtDNA damage [140]. The nitroso-redox balance may sensitize cardiac ryanodine receptor (RyR2) to induce ventricular arrhythmias. This reaction leads to an imbalance of Ca2+ and increased formation of ROS [141]. Modulating ion channels is one target for arrhythmia treatment and many anti-arrhythmic medications target these channels [106, 142]. Radiofrequency ablation is used to block abnormal conduction bundles and the origin of tachyarrhythmias. All of these sustain ion homeostasis in cardiomyocytes.

    Aortic aneurysms

    Marfan’s syndrome (MFS) is a systemic disease with a high incidence of aortic aneurysm and aortic dissection. These conditions have a high mortality in MFS. ROS produce endothelial dysfunction, switch smooth muscle cell phenotype, and cause extracellular matrix remodeling, leading to the progression of MFS [143]. NOX is one of the sources of ROS production. However, NOX has a different function in the process of aortic aneurysm [144]. The lack of NOX1 has a protective effect in aortic aneurysm [145] (Figure 3). However, a deficiency of NOX2 can contribute to the development of aortic aneurysm due to activation of inflammatory processes. A low expression of NOX4 offers potential protection in aortic aneurysm by ameliorating elastic fiber [146] (Figure 3). ROS also participate in cell death which contributes to aortic aneurysm [147]. Treatment of antioxidant stress may provide a potential option for preventing aortic aneurysm.

    Conclusions

    ROS are highly reactive molecules produced by a system of oxidases which have a great impact in the progression of cardiovascular diseases. The production of ROS disrupts the function of mitochondria and intracellular Ca2+ homeostasis, leading to damage to the cardiovascular system. Several mice models demonstrate that modulation of different pathways can rescue the impaired cardiomyocytes, retard myocardial remodeling, and maintain ion homeostasis. Therapies that target the activation of antioxidant systems, such as an exogenous antioxidant supplement, may be an effective treatment option in cardiovascular diseases. Further research is needed to explore the effect of controlling ROS in cardiovascular diseases.

    Abbreviations

    AF:

    atrial fibrillation

    AMPK:

    adenosine monophosphate-activated protein kinase

    ATP:

    adenosine triphosphate

    BH4:

    tetrahydrobiopterin

    Cx43:

    connexin 43

    DCM:

    diabetic cardiomyopathy

    eNOS:

    endothelial nitric oxide synthase

    ETC:

    electron transport chain

    Foxo3α:

    forkhead box transcription factor 3α

    GSH:

    glutathione

    H2O2:

    hydrogen peroxide

    HIF:

    hypoxia-inducible transcription factor

    I/R:

    ischemia-reperfusion injury

    iNOS:

    inducible nitric oxide synthase

    KLF2:

    Kruppel-like factor 2

    MAOs:

    monoamine oxidases

    MAPK:

    mitogen-activated protein kinases

    MFS:

    Marfan’s syndrome

    mPTP:

    mitochondrial permeability transition pore

    mtDNA:

    mitochondrial DNA

    mtROS:

    mitochondrial ROS

    NAC:

    N-acetylcysteine

    NAD+:

    nicotinamide adenine dinucleotide

    NADPH:

    nicotinamide adenine dinucleotide phosphate

    NF-κB:

    nuclear factor kappaB

    NO:

    nitric oxide

    NOS:

    nitric oxide synthase

    NOXs:

    nicotinamide adenine dinucleotide phosphate oxidases

    Nrf2:

    nuclear factor E2-related factor 2

    PKC:

    protein kinase C

    PKM2:

    pyruvate kinase M2

    Ripk3:

    receptor-interacting protein 3

    ROS:

    reactive oxygen species

    Sirt1:

    sirtuin 1

    SOD:

    superoxide dismutase

    STAT3:

    signal transducer and activator of transcription 3

    XO:

    xanthine oxidase

    Declarations

    Author contributions

    JF, LJD and JR contributed conception and design of the study; JF and JR wrote the first draft of the manuscript. All authors contributed to manuscript revision, read and approved the submitted version.

    Conflicts of interest

    The authors declare that they have no conflicts of interest.

    Ethical approval

    Not applicable.

    Consent to participate

    Not applicable.

    Consent to publication

    Not applicable.

    Availability of data and materials

    Not applicable.

    Funding

    Work in our lab has been supported by the National Key R&D Program of China (2017YFA0506000). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

    Copyright

    © The Author(s) 2022.

    References

    Heymes C, Bendall JK, Ratajczak P, Cave AC, Samuel JL, Hasenfuss G, et al. Increased myocardial NADPH oxidase activity in human heart failure. J Am Coll Cardiol. 2003;41:216471. [DOI] [PubMed]
    Aroor AR, Mandavia C, Ren J, Sowers JR, Pulakat L. Mitochondria and oxidative stress in the cardiorenal metabolic syndrome. Cardiorenal Med. 2012;2:87109. [DOI] [PubMed] [PMC]
    Wang S, Guo W, Ren J. Stress signaling in paraquat-induced target organ toxicity. React Oxygen Species. 2016;1:13140. [DOI]
    Kang Q, Yang C. Oxidative stress and diabetic retinopathy: molecular mechanisms, pathogenetic role and therapeutic implications. Redox Biol. 2020;37:101799. [DOI] [PubMed] [PMC]
    Dubois-Deruy E, Peugnet V, Turkieh A, Pinet F. Oxidative stress in cardiovascular diseases. Antioxidants. 2020;9:864. [DOI] [PubMed] [PMC]
    Martins-Marques T, Rodriguez-Sinovas A, Girao H. Cellular crosstalk in cardioprotection: where and when do reactive oxygen species play a role? Free Radic Biol Med. 2021;169:397409. [DOI] [PubMed]
    Sies H, Berndt C, Jones DP. Oxidative stress. Annu Rev Biochem. 2017;86:71548. [DOI] [PubMed]
    Liang T, Gao F, Chen J. Role of PTEN-less in cardiac injury, hypertrophy and regeneration. Cell Regen. 2021;10:25. [DOI] [PubMed] [PMC]
    Zhang W, Liang J, Han P. Cardiac cell type-specific responses to injury and contributions to heart regeneration. Cell Regen. 2021;10:4. [DOI] [PubMed] [PMC]
    Li H, Chang C, Li X, Zhang R. The roles and activation of endocardial Notch signaling in heart regeneration. Cell Regen. 2021;10:3. [DOI] [PubMed] [PMC]
    Guo R, Ma H, Gao F, Zhong L, Ren J. Metallothionein alleviates oxidative stress-induced endoplasmic reticulum stress and myocardial dysfunction. J Mol Cell Cardiol. 2009;47:22837. [DOI] [PubMed] [PMC]
    Lushchak VI. Free radicals, reactive oxygen species, oxidative stress and its classification. Chem Biol Interact. 2014;224:16475. [DOI] [PubMed]
    Münzel T, Gori T, Keaney JF Jr, Maack C, Daiber A. Pathophysiological role of oxidative stress in systolic and diastolic heart failure and its therapeutic implications. Eur Heart J. 2015;36:255564. [DOI] [PubMed] [PMC]
    Touyz RM, Rios FJ, Alves-Lopes R, Neves KB, Camargo LL, Montezano AC. Oxidative stress: a unifying paradigm in hypertension. Can J Cardiol. 2020;36:65970. [DOI] [PubMed] [PMC]
    Tang D, Chen X, Kang R, Kroemer G. Ferroptosis: molecular mechanisms and health implications. Cell Res. 2021;31:10725. [DOI] [PubMed] [PMC]
    Förstermann U, Xia N, Li H. Roles of vascular oxidative stress and nitric oxide in the pathogenesis of atherosclerosis. Circ Res. 2017;120:71335. [DOI] [PubMed]
    Förstermann U. Oxidative stress in vascular disease: causes, defense mechanisms and potential therapies. Nat Clin Pract Cardiovasc Med. 2008;5:33849. [DOI] [PubMed]
    Landmesser U, Spiekermann S, Preuss C, Sorrentino S, Fischer D, Manes C, et al. Angiotensin II induces endothelial xanthine oxidase activation: role for endothelial dysfunction in patients with coronary disease. Arterioscler Thromb Vasc Biol. 2007;27:9438. [DOI] [PubMed]
    Förstermann U. Nitric oxide and oxidative stress in vascular disease. Pflugers Arch. 2010;459:92339. [DOI] [PubMed]
    Kattoor AJ, Pothineni NVK, Palagiri D, Mehta JL. Oxidative stress in atherosclerosis. Curr Atheroscler Rep. 2017;19:42. [DOI] [PubMed]
    Marchio P, Guerra-Ojeda S, Vila JM, Aldasoro M, Victor VM, Mauricio MD. Targeting early atherosclerosis: a focus on oxidative stress and inflammation. Oxid Med Cell Longev. 2019;2019:8563845. [DOI] [PubMed] [PMC]
    Davidson SM, Duchen MR. Endothelial mitochondria: contributing to vascular function and disease. Circ Res. 2007;100:112841. [DOI] [PubMed]
    Kaludercic N, Di Lisa F. Mitochondrial ROS formation in the pathogenesis of diabetic cardiomyopathy. Front Cardiovasc Med. 2020;7:12. [DOI] [PubMed] [PMC]
    Zhou B, Tian R. Mitochondrial dysfunction in pathophysiology of heart failure. J Clin Invest. 2018;128:371626. [DOI] [PubMed] [PMC]
    Dedkova EN, Blatter LA. Modulation of mitochondrial Ca2+ by nitric oxide in cultured bovine vascular endothelial cells. Am J Physiol Cell Physiol. 2005;289:C83645. [DOI] [PubMed]
    Braunersreuther V, Montecucco F, Asrih M, Pelli G, Galan K, Frias M, et al. Role of NADPH oxidase isoforms NOX1, NOX2 and NOX4 in myocardial ischemia/reperfusion injury. J Mol Cell Cardiol. 2013;64:99107. [DOI] [PubMed]
    Hafstad AD, Nabeebaccus AA, Shah AM. Novel aspects of ROS signalling in heart failure. Basic Res Cardiol. 2013;108:359. [DOI] [PubMed]
    Zinkevich NS, Gutterman DD. ROS-induced ROS release in vascular biology: redox-redox signaling. Am J Physiol Heart Circ Physiol. 2011;301:H64753. [DOI] [PubMed] [PMC]
    Bedard K, Krause KH. The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology. Physiol Rev. 2007;87:245313. [DOI] [PubMed]
    Zhang Y, Murugesan P, Huang K, Cai H. NADPH oxidases and oxidase crosstalk in cardiovascular diseases: novel therapeutic targets. Nat Rev Cardiol. 2020;17:17094. [DOI] [PubMed] [PMC]
    Brandes RP, Schröder K. Differential vascular functions of Nox family NADPH oxidases. Curr Opin Lipidol. 2008;19:5138. [DOI] [PubMed]
    Ago T, Kitazono T, Kuroda J, Kumai Y, Kamouchi M, Ooboshi H, et al. NAD(P)H oxidases in rat basilar arterial endothelial cells. Stroke. 2005;36:10406. [DOI] [PubMed]
    Sirker A, Zhang M, Shah AM. NADPH oxidases in cardiovascular disease: insights from in vivo models and clinical studies. Basic Res Cardiol. 2011;106:73547. [DOI] [PubMed] [PMC]
    Gimenez M, Schickling BM, Lopes LR, Miller FJ Jr. Nox1 in cardiovascular diseases: regulation and pathophysiology. Clin Sci (Lond). 2016;130:15165. [DOI] [PubMed]
    Doughan AK, Harrison DG, Dikalov SI. Molecular mechanisms of angiotensin II-mediated mitochondrial dysfunction: linking mitochondrial oxidative damage and vascular endothelial dysfunction. Circ Res. 2008;102:48896. [DOI] [PubMed]
    van der Pol A, van Gilst WH, Voors AA, van der Meer P. Treating oxidative stress in heart failure: past, present and future. Eur J Heart Fail. 2019;21:42535. [DOI] [PubMed] [PMC]
    Privratsky JR, Wold LE, Sowers JR, Quinn MT, Ren J. AT1 blockade prevents glucose-induced cardiac dysfunction in ventricular myocytes: role of the AT1 receptor and NADPH oxidase. Hypertension. 2003;42:20612. [DOI] [PubMed]
    Canton M, Sánchez-Rodríguez R, Spera I, Venegas FC, Favia M, Viola A, et al. Reactive oxygen species in macrophages: sources and targets. Front Immunol. 2021;12:734229. [DOI] [PubMed] [PMC]
    Cappola TP, Kass DA, Nelson GS, Berger RD, Rosas GO, Kobeissi ZA, et al. Allopurinol improves myocardial efficiency in patients with idiopathic dilated cardiomyopathy. Circulation. 2001;104:240711. [DOI] [PubMed]
    Zhou Z, Ni K, Deng H, Chen X. Dancing with reactive oxygen species generation and elimination in nanotheranostics for disease treatment. Adv Drug Deliv Rev. 2020;158:7390. [DOI] [PubMed]
    Gong YY, Luo JY, Wang L, Huang Y. MicroRNAs regulating reactive oxygen species in cardiovascular diseases. Antioxid Redox Signal. 2018;29:1092107. [DOI] [PubMed]
    Ray PD, Huang BW, Tsuji Y. Reactive oxygen species (ROS) homeostasis and redox regulation in cellular signaling. Cell Signal. 2012;24:98190. [DOI] [PubMed] [PMC]
    Malinska D, Mirandola SR, Kunz WS. Mitochondrial potassium channels and reactive oxygen species. FEBS Lett. 2010;584:20438. [DOI] [PubMed]
    Hinkle PC, Butow RA, Racker E, Chance B. Partial resolution of the enzymes catalyzing oxidative phosphorylation. XV. Reverse electron transfer in the flavin-cytochrome beta region of the respiratory chain of beef heart submitochondrial particles. J Biol Chem. 1967;242:516973. [DOI] [PubMed]
    Sies H, Jones DP. Reactive oxygen species (ROS) as pleiotropic physiological signalling agents. Nat Rev Mol Cell Biol. 2020;21:36383. [DOI] [PubMed]
    Kimura S, Zhang GX, Nishiyama A, Shokoji T, Yao L, Fan YY, et al. Mitochondria-derived reactive oxygen species and vascular MAP kinases: comparison of angiotensin II and diazoxide. Hypertension. 2005;45:43844. [DOI] [PubMed]
    Han D, Williams E, Cadenas E. Mitochondrial respiratory chain-dependent generation of superoxide anion and its release into the intermembrane space. Biochem J. 2001;353:4116. [DOI] [PubMed] [PMC]
    Nurse CA, Salman S, Scott AL. Hypoxia-regulated catecholamine secretion in chromaffin cells. Cell Tissue Res. 2018;372:43341. [DOI] [PubMed]
    Read AD, Bentley RE, Archer SL, Dunham-Snary KJ. Mitochondrial iron-sulfur clusters: structure, function, and an emerging role in vascular biology. Redox Biol. 2021;47:102164. [DOI] [PubMed] [PMC]
    Okafor ON, Farrington K, Gorog DA. Allopurinol as a therapeutic option in cardiovascular disease. Pharmacol Ther. 2017;172:13950. [DOI] [PubMed]
    Kang SM, Lim S, Song H, Chang W, Lee S, Bae SM, et al. Allopurinol modulates reactive oxygen species generation and Ca2+ overload in ischemia-reperfused heart and hypoxia-reoxygenated cardiomyocytes. Eur J Pharmacol. 2006;535:2129. [DOI] [PubMed]
    Youdim MB, Edmondson D, Tipton KF. The therapeutic potential of monoamine oxidase inhibitors. Nat Rev Neurosci. 2006;7:295309. [DOI] [PubMed]
    Edmondson DE, Mattevi A, Binda C, Li M, Hubálek F. Structure and mechanism of monoamine oxidase. Curr Med Chem. 2004;11:198393. [DOI] [PubMed]
    Ramsay RR. Monoamine oxidases: the biochemistry of the proteins as targets in medicinal chemistry and drug discovery. Curr Top Med Chem. 2012;12:2189209. [DOI] [PubMed]
    Kaludercic N, Takimoto E, Nagayama T, Feng N, Lai EW, Bedja D, et al. Monoamine oxidase A-mediated enhanced catabolism of norepinephrine contributes to adverse remodeling and pump failure in hearts with pressure overload. Circ Res. 2010;106:193202. [DOI] [PubMed] [PMC]
    Bianchi P, Kunduzova O, Masini E, Cambon C, Bani D, Raimondi L, et al. Oxidative stress by monoamine oxidase mediates receptor-independent cardiomyocyte apoptosis by serotonin and postischemic myocardial injury. Circulation. 2005;112:3297305. [DOI] [PubMed]
    Pchejetski D, Kunduzova O, Dayon A, Calise D, Seguelas MH, Leducq N, et al. Oxidative stress-dependent sphingosine kinase-1 inhibition mediates monoamine oxidase A-associated cardiac cell apoptosis. Circ Res. 2007;100:419. [DOI] [PubMed]
    Förstermann U, Li H. Therapeutic effect of enhancing endothelial nitric oxide synthase (eNOS) expression and preventing eNOS uncoupling. Br J Pharmacol. 2011;164:21323. [DOI] [PubMed] [PMC]
    Faria A, Persaud SJ. Cardiac oxidative stress in diabetes: mechanisms and therapeutic potential. Pharmacol Ther. 2017;172:5062. [DOI] [PubMed]
    Eelen G, de Zeeuw P, Simons M, Carmeliet P. Endothelial cell metabolism in normal and diseased vasculature. Circ Res. 2015;116:123144. [DOI] [PubMed] [PMC]
    Comità S, Femmino S, Thairi C, Alloatti G, Boengler K, Pagliaro P, et al. Regulation of STAT3 and its role in cardioprotection by conditioning: focus on non-genomic roles targeting mitochondrial function. Basic Res Cardiol. 2021;116:56. [DOI] [PubMed] [PMC]
    Holmström KM, Finkel T. Cellular mechanisms and physiological consequences of redox-dependent signalling. Nat Rev Mol Cell Biol. 2014;15:41121. [DOI] [PubMed]
    Zhu P, Hu S, Jin Q, Li D, Tian F, Toan S, et al. Ripk3 promotes ER stress-induced necroptosis in cardiac IR injury: a mechanism involving calcium overload/XO/ROS/mPTP pathway. Redox Biol. 2018;16:15768. [DOI] [PubMed] [PMC]
    Le DD, Kim W, Lim S, Kim SC, Choi G. Identification of three groups of ginsenoside biosynthetic UDP-glycosyltransferases from Gynostemma pentaphyllum. Plant Sci. 2021;313:111069. [DOI] [PubMed]
    Duan J, Guan Y, Mu F, Guo C, Zhang E, Yin Y, et al. Protective effect of butin against ischemia/reperfusion-induced myocardial injury in diabetic mice: involvement of the AMPK/GSK-3β/Nrf2 signaling pathway. Sci Rep. 2017;7:41491. [DOI] [PubMed] [PMC]
    Zeng J, Jin Q, Ruan Y, Sun C, Xu G, Chu M, et al. Inhibition of TGFβ-activated protein kinase 1 ameliorates myocardial ischaemia/reperfusion injury via endoplasmic reticulum stress suppression. J Cell Mol Med. 2020;24:684659. [DOI] [PubMed] [PMC]
    Boengler K, Hilfiker-Kleiner D, Heusch G, Schulz R. Inhibition of permeability transition pore opening by mitochondrial STAT3 and its role in myocardial ischemia/reperfusion. Basic Res Cardiol. 2010;105:77185. [DOI] [PubMed] [PMC]
    Kunisada K, Negoro S, Tone E, Funamoto M, Osugi T, Yamada S, et al. Signal transducer and activator of transcription 3 in the heart transduces not only a hypertrophic signal but a protective signal against doxorubicin-induced cardiomyopathy. Proc Natl Acad Sci U S A. 2000;97:3159. [DOI] [PubMed] [PMC]
    Obana M, Maeda M, Takeda K, Hayama A, Mohri T, Yamashita T, et al. Therapeutic activation of signal transducer and activator of transcription 3 by interleukin-11 ameliorates cardiac fibrosis after myocardial infarction. Circulation. 2010;121:68491. [DOI] [PubMed]
    Yang H, Yang T, Baur JA, Perez E, Matsui T, Carmona JJ, et al. Nutrient-sensitive mitochondrial NAD+ levels dictate cell survival. Cell. 2007;130:1095107. [DOI] [PubMed] [PMC]
    Hsu CP, Oka S, Shao D, Hariharan N, Sadoshima J. Nicotinamide phosphoribosyltransferase regulates cell survival through NAD+ synthesis in cardiac myocytes. Circ Res. 2009;105:48191. [DOI] [PubMed] [PMC]
    Hund TJ, Lerner DL, Yamada KA, Schuessler RB, Saffitz JE. Protein kinase Cepsilon mediates salutary effects on electrical coupling induced by ischemic preconditioning. Heart Rhythm. 2007;4:118393. [DOI] [PubMed] [PMC]
    Hirschhäuser C, Lissoni A, Görge PM, Lampe PD, Heger J, Schlüter KD, et al. Connexin 43 phosphorylation by casein kinase 1 is essential for the cardioprotection by ischemic preconditioning. Basic Res Cardiol. 2021;116:21. [DOI] [PubMed] [PMC]
    Basheer WA, Fu Y, Shimura D, Xiao S, Agvanian S, Hernandez DM, et al. Stress response protein GJA1-20k promotes mitochondrial biogenesis, metabolic quiescence, and cardioprotection against ischemia/reperfusion injury. JCI Insight. 2018;3:e121900. [DOI] [PubMed] [PMC]
    Pecoraro M, Pinto A, Popolo A. Inhibition of connexin 43 translocation on mitochondria accelerates CoCl2-induced apoptotic response in a chemical model of hypoxia. Toxicol In Vitro. 2018;47:1208. [DOI] [PubMed]
    Chouchani ET, Pell VR, Gaude E, Aksentijević D, Sundier SY, Robb EL, et al. Ischaemic accumulation of succinate controls reperfusion injury through mitochondrial ROS. Nature. 2014;515:4315. [DOI] [PubMed] [PMC]
    Heusch G. Myocardial stunning and hibernation revisited. Nat Rev Cardiol. 2021;18:52236. [DOI] [PubMed]
    Ushio-Fukai M, Ash D, Nagarkoti S, Belin de Chantemèle EJ, Fulton DJR, Fukai T. Interplay between reactive oxygen/reactive nitrogen species and metabolism in vascular biology and disease. Antioxid Redox Signal. 2021;34:131954. [DOI] [PubMed] [PMC]
    Janbandhu V, Tallapragada V, Patrick R, Li Y, Abeygunawardena D, Humphreys DT, et al. Hif-1a suppresses ROS-induced proliferation of cardiac fibroblasts following myocardial infarction. Cell Stem Cell. 2022;29:28197.e12. [DOI] [PubMed] [PMC]
    Shimura D, Nuebel E, Baum R, Valdez SE, Xiao S, Warren JS, et al. Protective mitochondrial fission induced by stress-responsive protein GJA1-20k. Elife. 2021;10:e69207. [DOI] [PubMed] [PMC]
    Guo Y, Jia P, Chen Y, Yu H, Xin X, Bao Y, et al. PHLDA1 is a new therapeutic target of oxidative stress and ischemia reperfusion-induced myocardial injury. Life Sci. 2020;245:117347. [DOI] [PubMed]
    Korge P, Ping P, Weiss JN. Reactive oxygen species production in energized cardiac mitochondria during hypoxia/reoxygenation: modulation by nitric oxide. Circ Res. 2008;103:87380. [DOI] [PubMed] [PMC]
    Eble JA, Niland S. The extracellular matrix of blood vessels. Curr Pharm Des. 2009;15:1385400. [DOI] [PubMed]
    Hong FF, Liang XY, Liu W, Lv S, He SJ, Kuang HB, et al. Roles of eNOS in atherosclerosis treatment. Inflamm Res. 2019;68:42941. [DOI] [PubMed]
    Li H, Junk P, Huwiler A, Burkhardt C, Wallerath T, Pfeilschifter J, et al. Dual effect of ceramide on human endothelial cells: induction of oxidative stress and transcriptional upregulation of endothelial nitric oxide synthase. Circulation. 2002;106:22506. [DOI] [PubMed]
    Wu B, You S, Qian H, Wu S, Lu S, Zhang Y, et al. The role of SIRT2 in vascular-related and heart-related diseases: a review. J Cell Mol Med. 2021;25:64708. [DOI] [PubMed] [PMC]
    Tiyerili V, Zimmer S, Jung S, Wassmann K, Naehle CP, Lütjohann D, et al. CB1 receptor inhibition leads to decreased vascular AT1 receptor expression, inhibition of oxidative stress and improved endothelial function. Basic Res Cardiol. 2010;105:46577. [DOI] [PubMed]
    Tseng AH, Shieh SS, Wang DL. SIRT3 deacetylates FOXO3 to protect mitochondria against oxidative damage. Free Radic Biol Med. 2013;63:22234. [DOI] [PubMed]
    Bergaggio E, Riganti C, Garaffo G, Vitale N, Mereu E, Bandini C, et al. IDH2 inhibition enhances proteasome inhibitor responsiveness in hematological malignancies. Blood. 2019;133:15667. [DOI] [PubMed]
    Porter GA, Urciuoli WR, Brookes PS, Nadtochiy SM. SIRT3 deficiency exacerbates ischemia-reperfusion injury: implication for aged hearts. Am J Physiol Heart Circ Physiol. 2014;306:H16029. [DOI] [PubMed] [PMC]
    Jain T, Nikolopoulou EA, Xu Q, Qu A. Hypoxia inducible factor as a therapeutic target for atherosclerosis. Pharmacol Ther. 2018;183:2233. [DOI] [PubMed]
    Kim B, Jang C, Dharaneeswaran H, Li J, Bhide M, Yang S, et al. Endothelial pyruvate kinase M2 maintains vascular integrity. J Clin Invest. 2018;128:454356. [DOI] [PubMed] [PMC]
    Wu X, Liu L, Zheng Q, Hao H, Ye H, Li P, et al. Protocatechuic aldehyde protects cardiomycoytes against ischemic injury via regulation of nuclear pyruvate kinase M2. Acta Pharm Sin B. 2021;11:355366. [DOI] [PubMed] [PMC]
    Cheng J, Luo X, Huang Z, Chen L. Apelin/APJ system: a potential therapeutic target for endothelial dysfunction-related diseases. J Cell Physiol. 2019;234:1214960. [DOI] [PubMed]
    Azizi Y, Faghihi M, Imani A, Roghani M, Zekri A, Mobasheri MB, et al. Post-infarct treatment with [Pyr(1)] apelin-13 improves myocardial function by increasing neovascularization and overexpression of angiogenic growth factors in rats. Eur J Pharmacol. 2015;761:1018. [DOI] [PubMed]
    Zhao H, Yao P, Li L, Chen L. Apelin receptor signaling: a novel mechanism of endothelial cell polarization. Acta Biochim Biophys Sin (Shanghai). 2016;48:11389. [DOI] [PubMed]
    Leisegang MS, Bibli SI, Günther S, Pflüger-Müller B, Oo JA, Höper C, et al. Pleiotropic effects of laminar flow and statins depend on the Krüppel-like factor-induced lncRNA MANTIS. Eur Heart J. 2019;40:252333. [DOI] [PubMed]
    Emini Veseli B, Perrotta P, De Meyer GRA, Roth L, Van der Donckt C, Martinet W, et al. Animal models of atherosclerosis. Eur J Pharmacol. 2017;816:313. [DOI] [PubMed]
    Nakano H, Nakajima A, Sakon-Komazawa S, Piao JH, Xue X, Okumura K. Reactive oxygen species mediate crosstalk between NF-kappaB and JNK. Cell Death Differ. 2006;13:7307. [DOI] [PubMed]
    Liu J, Yi X, Tao Y, Wang Y, Xu Z. Insulin-receptor substrate 1 protects against injury in endothelial cell models of ox-LDL-induced atherosclerosis by inhibiting ER stress/oxidative stress-mediated apoptosis and activating the Akt/FoxO1 signaling pathway. Int J Mol Med. 2020;46:167182. [DOI] [PubMed] [PMC]
    Paneni F, Beckman JA, Creager MA, Cosentino F. Diabetes and vascular disease: pathophysiology, clinical consequences, and medical therapy: part I. Eur Heart J. 2013;34:243643. [DOI] [PubMed] [PMC]
    Horke S, Witte I, Wilgenbus P, Krüger M, Strand D, Förstermann U. Paraoxonase-2 reduces oxidative stress in vascular cells and decreases endoplasmic reticulum stress-induced caspase activation. Circulation. 2007;115:205564. [DOI] [PubMed]
    Karunakaran D, Nguyen MA, Geoffrion M, Vreeken D, Lister Z, Cheng HS, et al. RIPK1 expression associates with inflammation in early atherosclerosis in humans and can be therapeutically silenced to reduce NF-κB activation and atherogenesis in mice. Circulation. 2021;143:16377. [DOI] [PubMed]
    Lv K, Kong L, Yang M, Zhang L, Chu S, Zhang L, et al. An apoA-I mimic peptide of 4F promotes SDF-1α expression in endothelial cells through PI3K/Akt/ERK/HIF-1α signaling pathway. Front Pharmacol. 2022;12:760908. [DOI] [PubMed] [PMC]
    Li Y, Sun R, Zou J, Ying Y, Luo Z. Dual roles of the AMP-activated protein kinase pathway in angiogenesis. Cells. 2019;8:752. [DOI] [PubMed] [PMC]
    Touyz RM, Anagnostopoulou A, Camargo LL, Rios FJ, Montezano AC. Vascular biology of superoxide-generating NADPH oxidase 5-implications in hypertension and cardiovascular disease. Antioxid Redox Signal. 2019;30:102740. [DOI] [PubMed] [PMC]
    Takimoto E, Champion HC, Li M, Ren S, Rodriguez ER, Tavazzi B, et al. Oxidant stress from nitric oxide synthase-3 uncoupling stimulates cardiac pathologic remodeling from chronic pressure load. J Clin Invest. 2005;115:122131. [DOI] [PubMed] [PMC]
    Kröller-Schön S, Jansen T, Tran TLP, Kvandová M, Kalinovic S, Oelze M, et al. Endothelial α1AMPK modulates angiotensin II-mediated vascular inflammation and dysfunction. Basic Res Cardiol. 2019;114:8. [DOI] [PubMed]
    Liu Y, Rafferty TM, Rhee SW, Webber JS, Song L, Ko B, et al. CD8+ T cells stimulate Na-Cl co-transporter NCC in distal convoluted tubules leading to salt-sensitive hypertension. Nat Commun. 2017;8:14037. [DOI] [PubMed] [PMC]
    Dikalova AE, Pandey A, Xiao L, Arslanbaeva L, Sidorova T, Lopez MG, 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;126:43952. [DOI] [PubMed] [PMC]
    Dolinsky VW, Chakrabarti S, Pereira TJ, Oka T, Levasseur J, Beker D, et al. Resveratrol prevents hypertension and cardiac hypertrophy in hypertensive rats and mice. Biochim Biophys Acta. 2013;1832:172333. [DOI] [PubMed]
    Hu X, Wu P, Liu B, Lang Y, Li T. RNA-binding protein CELF1 promotes cardiac hypertrophy via interaction with PEBP1 in cardiomyocytes. Cell Tissue Res. 2022;387:11121. [DOI] [PubMed]
    Cates C, Rousselle T, Wang J, Quan N, Wang L, Chen X, et al. Activated protein C protects against pressure overload-induced hypertrophy through AMPK signaling. Biochem Biophys Res Commun. 2018;495:258494. [DOI] [PubMed] [PMC]
    Wang B, Nie J, Wu L, Hu Y, Wen Z, Dong L, et al. AMPKα2 protects against the development of heart failure by enhancing mitophagy via PINK1 phosphorylation. Circ Res. 2018;122:71229. [DOI] [PubMed] [PMC]
    Gélinas R, Mailleux F, Dontaine J, Bultot L, Demeulder B, Ginion A, et al. AMPK activation counteracts cardiac hypertrophy by reducing O-GlcNAcylation. Nat Commun. 2018;9:374. [DOI] [PubMed] [PMC]
    Ortega MA, Poirion O, Zhu X, Huang S, Wolfgruber TK, Sebra R, et al. Using single-cell multiple omics approaches to resolve tumor heterogeneity. Clin Transl Med. 2017;6:46. [DOI] [PubMed] [PMC]
    Zang R, Tan Q, Zeng F, Wang D, Yu S, Wang Q. JMJD1A represses the development of cardiomyocyte hypertrophy by regulating the expression of Catalase. Biomed Res Int. 2020;2020:5081323. [DOI] [PubMed] [PMC]
    Li S, Zhu Z, Xue M, Yi X, Liang J, Niu C, et al. Fibroblast growth factor 21 protects the heart from angiotensin II-induced cardiac hypertrophy and dysfunction via SIRT1. Biochim Biophys Acta Mol Basis Dis. 2019;1865:124152. [DOI] [PubMed]
    Zhang M, Brewer AC, Schröder K, Santos CX, Grieve DJ, Wang M, et al. NADPH oxidase-4 mediates protection against chronic load-induced stress in mouse hearts by enhancing angiogenesis. Proc Natl Acad Sci U S A. 2010;107:181216. [DOI] [PubMed] [PMC]
    Montezano AC, Dulak-Lis M, Tsiropoulou S, Harvey A, Briones AM, Touyz RM. Oxidative stress and human hypertension: vascular mechanisms, biomarkers, and novel therapies. Can J Cardiol. 2015;31:63141. [DOI] [PubMed]
    Dovinova I, Kvandová M, Balis P, Gresova L, Majzunova M, Horakova L, et al. The role of Nrf2 and PPARgamma in the improvement of oxidative stress in hypertension and cardiovascular diseases. Physiol Res. 2020;69:S54153. [DOI] [PubMed] [PMC]
    Jarosz J, Ghosh S, Delbridge LM, Petzer A, Hickey AJ, Crampin EJ, et al. Changes in mitochondrial morphology and organization can enhance energy supply from mitochondrial oxidative phosphorylation in diabetic cardiomyopathy. Am J Physiol Cell Physiol. 2017;312:C1907. [DOI] [PubMed]
    Wold LE, Ceylan-Isik AF, Ren J. Oxidative stress and stress signaling: menace of diabetic cardiomyopathy. Acta Pharmacol Sin. 2005;26:90817. [DOI] [PubMed]
    Bernardi P, Di Lisa F, Fogolari F, Lippe G. From ATP to PTP and back: a dual function for the mitochondrial ATP synthase. Circ Res. 2015;116:185062. [DOI] [PubMed] [PMC]
    Sivitz WI, Yorek MA. Mitochondrial dysfunction in diabetes: from molecular mechanisms to functional significance and therapeutic opportunities. Antioxid Redox Signal. 2010;12:53777. [DOI] [PubMed] [PMC]
    Wu S, Lu Q, Ding Y, Wu Y, Qiu Y, Wang P, et al. Hyperglycemia-driven inhibition of AMP-activated protein kinase α2 induces diabetic cardiomyopathy by promoting mitochondria-associated endoplasmic reticulum membranes in vivo. Circulation. 2019;139:191336. [DOI] [PubMed] [PMC]
    Wei H, Bu R, Yang Q, Jia J, Li T, Wang Q, et al. Exendin-4 protects against hyperglycemia-induced cardiomyocyte pyroptosis via the AMPK-TXNIP pathway. J Diabetes Res. 2019;2019:8905917. [DOI] [PubMed] [PMC]
    Pelletier A, Coderre L. Ketone bodies alter dinitrophenol-induced glucose uptake through AMPK inhibition and oxidative stress generation in adult cardiomyocytes. Am J Physiol Endocrinol Metab. 2007;292:E132532. [DOI] [PubMed]
    Guo Z, Tuo H, Tang N, Liu FY, Ma SQ, An P, et al. Neuraminidase 1 deficiency attenuates cardiac dysfunction, oxidative stress, fibrosis, inflammatory via AMPK-SIRT3 pathway in diabetic cardiomyopathy mice. Int J Biol Sci. 2022;18:82640. [DOI] [PubMed] [PMC]
    Hansen SS, Aasum E, Hafstad AD. The role of NADPH oxidases in diabetic cardiomyopathy. Biochim Biophys Acta Mol Basis Dis. 2018;1864:190813. [DOI] [PubMed]
    Maalouf RM, Eid AA, Gorin YC, Block K, Escobar GP, Bailey S, et al. Nox4-derived reactive oxygen species mediate cardiomyocyte injury in early type 1 diabetes. Am J Physiol Cell Physiol. 2012;302:C597604. [DOI] [PubMed] [PMC]
    Bertero E, Maack C. Calcium signaling and reactive oxygen species in mitochondria. Circ Res. 2018;122:146078. [DOI] [PubMed]
    Wagner S, Rokita AG, Anderson ME, Maier LS. Redox regulation of sodium and calcium handling. Antioxid Redox Signal. 2013;18:106377. [DOI] [PubMed] [PMC]
    Yamaguchi O, Higuchi Y, Hirotani S, Kashiwase K, Nakayama H, Hikoso S, et al. Targeted deletion of apoptosis signal-regulating kinase 1 attenuates left ventricular remodeling. Proc Natl Acad Sci U S A. 2003;100:158838. [DOI] [PubMed] [PMC]
    Wang C, Wang Y, Shen L. Mitochondrial proteins in heart failure: the role of deacetylation by SIRT3. Pharmacol Res. 2021;172:105802. [DOI] [PubMed]
    Karamanlidis G, Lee CF, Garcia-Menendez L, Kolwicz SC Jr, Suthammarak W, Gong G, et al. Mitochondrial complex I deficiency increases protein acetylation and accelerates heart failure. Cell Metab. 2013;18:23950. [DOI] [PubMed] [PMC]
    Chen J, Chen S, Zhang B, Liu J. SIRT3 as a potential therapeutic target for heart failure. Pharmacol Res. 2021;165:105432. [DOI] [PubMed]
    Mason FE, Pronto JRD, Alhussini K, Maack C, Voigt N. Cellular and mitochondrial mechanisms of atrial fibrillation. Basic Res Cardiol. 2020;115:72. [DOI] [PubMed] [PMC]
    Friedrichs K, Baldus S, Klinke A. Fibrosis in atrial fibrillation—role of reactive species and MPO. Front Physiol. 2012;3:214. [DOI] [PubMed] [PMC]
    Anderson EJ, Efird JT, Davies SW, O’Neal WT, Darden TM, Thayne KA, et al. Monoamine oxidase is a major determinant of redox balance in human atrial myocardium and is associated with postoperative atrial fibrillation. J Am Heart Assoc. 2014;3:e000713. [DOI] [PubMed] [PMC]
    Cutler MJ, Plummer BN, Wan X, Sun QA, Hess D, Liu H, et al. Aberrant S-nitrosylation mediates calcium-triggered ventricular arrhythmia in the intact heart. Proc Natl Acad Sci U S A. 2012;109:1818691. [DOI] [PubMed] [PMC]
    Jeong EM, Liu M, Sturdy M, Gao G, Varghese ST, Sovari AA, et al. Metabolic stress, reactive oxygen species, and arrhythmia. J Mol Cell Cardiol. 2012;52:45463. [DOI] [PubMed] [PMC]
    Rysz J, Gluba-Brzózka A, Rokicki R, Franczyk B. Oxidative stress-related susceptibility to aneurysm in Marfan’s syndrome. Biomedicines. 2021;9:1171. [DOI] [PubMed] [PMC]
    Kigawa Y, Miyazaki T, Lei XF, Kim-Kaneyama JR, Miyazaki A. Functional heterogeneity of Nadph oxidases in atherosclerotic and aneurysmal diseases. J Atheroscler Thromb. 2017;24:113. [DOI] [PubMed] [PMC]
    Gavazzi G, Deffert C, Trocme C, Schäppi M, Herrmann FR, Krause KH. NOX1 deficiency protects from aortic dissection in response to angiotensin II. Hypertension. 2007;50:18996. [DOI] [PubMed]
    Jiménez-Altayó F, Meirelles T, Crosas-Molist E, Sorolla MA, Del Blanco DG, López-Luque J, et al. Redox stress in Marfan syndrome: dissecting the role of the NADPH oxidase NOX4 in aortic aneurysm. Free Radic Biol Med. 2018;118:4458. [DOI] [PubMed]
    Adiguzel Z, Arda N, Kacar O, Serhatli M, Gezer Tas S, Baykal AT, et al. Evaluation of apoptotic molecular pathways for smooth muscle cells isolated from thoracic aortic aneurysms in response to oxidized sterols. Mol Biol Rep. 2014;41:787584. [DOI] [PubMed]