Next Article in Journal
HER2 Low Breast Cancer: A New Subtype or a Trojan for Cytotoxic Drug Delivery?
Previous Article in Journal
Populus euphratica GLABRA3 Binds PLDδ Promoters to Enhance Salt Tolerance
Previous Article in Special Issue
Cardiac-Specific Overexpression of ERRγ in Mice Induces Severe Heart Dysfunction and Early Lethality
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mitochondrial Dysfunction in the Cardio-Renal Axis

1
Laboratory of Vascular Pathology, IIS-Fundación Jiménez Díaz, 28040 Madrid, Spain
2
Centro de Investigación Biomédica en Red de Enfermedades Cardiovasculares (CIBERCV), Faculty of Medicine and Biomedicine, Universidad Alfonso X El Sabio, 28037 Madrid, Spain
3
Spain Nephrology Laboratory, IIS-Fundación Jiménez Díaz-Universidad Autónoma, 28040 Madrid, Spain
4
REDINREN Spain/Ricors2040, 28029 Madrid, Spain
5
Cellular Biology in Renal Diseases Laboratory, IIS-Fundación Jiménez Díaz-Universidad Autónoma, 28040 Madrid, Spain
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(9), 8209; https://doi.org/10.3390/ijms24098209
Submission received: 29 March 2023 / Revised: 26 April 2023 / Accepted: 28 April 2023 / Published: 3 May 2023
(This article belongs to the Special Issue Cardiometabolic and Cardiotoxicity Disorders)

Abstract

:
Cardiovascular disease (CVD) frequently complicates chronic kidney disease (CKD). The risk of all-cause mortality increases from 20% to 500% in patients who suffer both conditions; this is referred to as the so-called cardio-renal syndrome (CRS). Preclinical studies have described the key role of mitochondrial dysfunction in cardiovascular and renal diseases, suggesting that maintaining mitochondrial homeostasis is a promising therapeutic strategy for CRS. In this review, we explore the malfunction of mitochondrial homeostasis (mitochondrial biogenesis, dynamics, oxidative stress, and mitophagy) and how it contributes to the development and progression of the main vascular pathologies that could be affected by kidney injury and vice versa, and how this knowledge may guide the development of novel therapeutic strategies in CRS.

1. Introduction

The incidence of heart failure and chronic kidney disease (CKD) is increasing worldwide [1]. The cardio-renal syndrome (CRS) was described in 2010 by the Acute Dialysis Quality Initiative (ADQI) as a heart and kidney disorder that is produced by an acute or chronic dysfunction in any of these organs that may induce a dysfunction of the other [2,3]. The global prevalence of CKD is estimated at 850 million people, of whom 10–47% had cardiovascular disease (CVD) [4,5]. CVD is the main cause of death in the CKD population [4]. In patients with heart failure, kidney failure increased the risk of all-cause mortality from 20% to 500% [6,7]. Indeed, CRS is identified as a public health burden due to its poorly understood pathogenesis and suboptimal treatment possibilities [8]. Currently, there are five types of CRS: (1) type-I CRS (acute cardiac failure that induces acute renal failure); (2) type-II CRS (chronic cardiac failure that leads chronic renal failure); (3) type-III CRS (acute kidney injury aggravating heart failure); (4) type-IV CRS (chronic kidney failure aggravating heart failure); and (5) type-V CRS (concurrent chronic cardiac and renal failure) [9]. However, due to the current controversy around these classifications, an updated CRS classification needs to be established.
Mitochondrial dysfunction is thought to play a key role in cardiovascular and kidney disease due to the high content of mitochondria and oxygen consumption in these tissues [10,11]. Mitochondria rapidly respond to insults and try to maintain homeostasis through changes in morphology and bioenergetics and upregulating self-renewal/degradation, and could be a bridge player, connecting the kidneys and heart in CRS [12,13]. Altered mitochondrial homeostasis that begins in the kidney as well as in the cardiovascular system may contribute to impaired cardiovascular and renal function independently or both at the same time through metabolism changes, chronic inflammation, oxidative stress, calcium dynamics, fibrosis, and autophagy [14,15,16].
Mitochondrial homeostasis changes could be not only a consequence of cardio-renal diseases, but also a possible common denominator between the kidney and cardiovascular system in the development of CRS. However, the vast majority of the studies about this field are focused on one pathology or another, but not how one affects the other (CRS). Currently, there are pathological elements targeting mitochondria that can commonly affect both renal and cardiovascular tissues simultaneously, such as circulating uremic toxins or hypertension. Studies show how the presence of uremic toxins resulting from renal malfunction can induce mitochondrial damage at the cardiovascular level [17]. On the other hand, hypertension is a common pathological event in CVD and CKD patients. Renal mitochondrial oxidative stress has been associated with renin–angiotensin system activation [18], leading to increased Angiotensin II and the subsequent induction of oxidative stress and, therefore, mitochondrial dysfunction in cardiovascular tissues, participating in the progression of damage [19]. These common pathological events demonstrate the close relationship between CKD–CVD and the important role of mitochondria in the context of CRS.
In this article, we summarize the role of mitochondrial homeostasis dysfunction in processes such as biogenesis, dynamics, oxidative stress, and mitophagy in the main vascular pathologies that could be affected by kidney injury and vice versa, and discuss novel therapeutic strategies for CRS. At the moment, more studies focused on the communication of the cardiovascular and kidney organ are needed to elucidate its crosstalk during pathological situations.

2. Mitochondrial Function under Physiological Conditions

Mitochondria are essential organelles whose main function is ATP production. However, lot of works have pointed out the functions of these organelles in processes such as cell death, intracellular calcium signaling, inflammatory response, metal ion homeostasis, redox state maintenance, phospholipid trafficking, and biosynthetic pathways. As such, the number and functionality of mitochondria is key to cell integrity and viability. In response to cell demands, mitochondria can change their morphology and number through mitochondrial biogenesis to generate new mitochondria and through mitochondrial dynamics involving fusion and fission. In mitochondrial fusion, two neighboring mitochondria join. This helps to keep them healthy and increases the capacity for oxidative phosphorylation during energy starvation [20,21]. Mitochondrial fission helps to remove damaged mitochondria by mitophagy and is also necessary for redistributing mitochondrial contents to daughter cells during mitosis [21].
Heart cells together with vessel wall cells conform to a high metabolic consumption and one of the most mitochondria abundant tissues in the human body. The kidney is the second organ for mitochondrial abundance since tubular cells require large amounts of energy to maintain their reabsorption function. In both the healthy kidneys and heart, fatty acid oxidation (FAO) and mitochondrial oxidative phosphorylation (OXPHOS) are the preferred pathway to produce ATP, with glycolysis being a minor pathway [10,22]. During acute kidney injury (AKI), tubular cells undergo metabolic reprogramming, as FAO genes are downregulated while glycolytic enzyme gene expression increases to restore the energy supply. However, FAO dysfunction leads to ATP depletion, lipid accumulation, inflammation, and fibrosis [23]. FAO is also modified in endothelial dysfunction [24] and decreased in heart failure; however, in this case, the alternative source of ATP is unclear [25].
Mitochondrial antioxidant defences are crucial for scavenging reactive oxygen species (ROS) generated by the electron transport chain (ETC) and they include enzymatic and non-enzymatic systems [26]. Among the enzymatic systems, superoxidase dismutase (SOD) converts the most dangerous ROS, superoxide anion (•O2), to hydrogen peroxide (H2O2), and H2O2 is reduced to water by mitochondrial catalase and glutathione peroxidases (a major scavenger in mitochondria to deal with H2O2 reduction) [27,28]. In this line, physiological low levels of mitochondrial reactive oxygen species (mtROS) can activate survival pathways, whereas excess mtROS cause oxidative stress and eventually cell death, so a balance between mtROS production and scavenging is essential to maintain healthy mitochondria and cells [29,30,31].

3. Mitochondrial Biogenesis

Mitochondrial biogenesis is the cellular process by which new mitochondria are produced [32]. However, mitochondria cannot be produced de novo from scratch; they are produced by adding new contents (proteins and membranes) to pre-existing mitochondria, following a fission process similar to prokaryotic binary fission [33]. Mitochondrial biogenesis is a complex process that involves both the mitochondrial and nuclear-encoded proteins and replication of mtDNA, requiring a co-ordinated regulation of both genomes by the transcriptional coactivator peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC-1α) [34,35]. PGC-1α expression is altered in kidney and cardiac disease [35,36,37] and, in some cases, exerts its effects through direct interaction/coactivation with PPARs, ERRs, and NRF-1/NRF-2, among other transcription factors [38,39].

3.1. Role in Renal Damage

There is considerable evidence that mitochondrion biogenesis is altered in AKI, and this negatively affects kidney function. Healthy proximal tubular cells are rich in mitochondria and express high levels of PGC-1α whose expression is severely reduced in human AKI and in preclinical AKI induced by lipopolysaccharide (LPS), ischemia reperfusion injury (IRI), folic acid overdose, or cisplatin overdose [40,41,42,43,44]. PGC-1α was the transcriptional regulator with the most severely decreased activity in folic acid AKI [44]. In both LPS and folic acid AKI, the reduced PGC-1α expression correlated with the reduced expression of its mitochondrial downstream target genes and negatively correlated with the serum urea levels, suggesting that PGC-1α downregulation during AKI has functional consequences. This was confirmed in PGC-1α knockout mice that showed persistent AKI following saline resuscitation LPS and more severe AKI following folic acid or IRI than wild-type mice [40,41,45]. In this line, the reduced renal mitochondrial mass observed in folic acid AKI was more severe in PGC-1α-deficient mice and negatively correlated with plasma creatinine levels [44]. Additionally, the transgenic overexpression of PGC-1α or treatment with drugs (NAD+ precursor Nicotinamide (NAM), AMPK, and Phosphodiesterase inhibitors and anti-TWEAK antibodies) that preserve PGC-1α expression resulted in increased mitochondrial mass and reduced kidney injury in experimental AKI [34,45], suggesting that therapeutic approaches that promote PGC-1α expression and/or mitochondrial mass may be beneficial in AKI.
There is a bidirectional relationship between inflammation and altered mitochondrial biogenesis. Inflammation may directly downregulate PGC-1α, as observed in cultured tubular cells exposed to the inflammatory cytokines TNFα or TWEAK, and in mice injected with TWEAK [40,41]. (TWEAK) is inflammatory cytokines member of the TNF superfamily that signals through its receptor FN14 and is involved in several biological processes, including proliferation, angiogenesis, induction of inflammatory cytokines, and, under some experimental conditions, apoptosis [46,47,48,49]. On the other hand, PGC-1α knockout mice develop spontaneous kidney inflammation as well as more severe kidney inflammatory responses during AKI than wild-type mice [44], supporting the existence of a positive feedback loop between inflammation and PGC-1α downregulation during AKI. Indeed, anti-inflammatory therapies may preserve PGC-1α expression in AKI. Thus, treatment with anti-TWEAK antibodies in folic acid AKI preserved kidney PGC1α levels, reduced kidney inflammation, and improved kidney function [40,50].
There is also evidence of altered mitochondrial biogenesis in CKD. Both a metabolomic study in urine from patients with diabetic CKD and a transcriptomic study in microdissected kidney tubule samples from CKD patients of different etiologies showed that mitochondrial biogenesis, FAO genes, and PGC-1α expression were impaired in CKD patients compared to healthy controls [51,52,53]. Furthermore, PGC-1α expression is also reduced in animal models of kidney fibrosis induced by folic acid overdose or by Neurogenic locus notch homolog protein 1 (Notch) overexpression, and tubule-specific overexpression of PGC-1α ameliorated Notch-induced kidney fibrosis [53]. The expression of PGC-1α is also reduced in podocytes of diabetic kidney disease (DKD) patients and in animal models of DKD, but the specific overexpression of PGC-1α in podocytes altered the mitochondrial properties and caused albuminuria and glomerulosclerosis [54]. In addition, a study described the key role of other downstream PGC-1α transcription factors in mitochondrial biogenesis [38,39]. A study in renal cancer cells showed that Mitochondrial pyruvate carrier (MPC) 1 is a novel target gene of PGC-1α. In this study, we demonstrated that the reduced expression of PGC-1α diminished the MPC expression, triggering the impaired mitochondrial respiratory capacity, reducing pyruvate transport into the mitochondrial matrix [55]. In renal cells from TFAM KO mice, we can see the induced cytosolic translocation and activation of the cytosolic cGAS-stimulator of interferon genes (STING) that induced the aberrant packaging of the mitochondrial DNA (mtDNA), cytokine expression, and immune cell recruitment [56]. These findings suggest a critical window of PGC-1α activity for AKI and CKD in regulating renal inflammation and fibrosis (Figure 1).

3.2. Role in Cardiovascular Damage

In CVD, mitochondrial biogenesis is decreased as characterized by a reduced mtDNA copy number or reduced metabolic enzymes associated with a decreased ATP production and/or enhanced ROS formation [57]. The main transcriptional activators linked to mitochondrial biogenesis belongs to the peroxisome proliferator-activated receptor γ-coactivator-1 family that includes PGC-1α and PGC-1β and PGC-1-related coactivator (PRC) [58,59]. Overexpression of PGC-1α in mice exacerbated mitochondrial biogenic responses associated with increased nuclear-encoded mitochondrial genes [60]. Germline targeting of PGC-1α and PGC-1β blocked mitochondrial biogenesis in the heart and triggered perinatal lethal heart failure [61].
In atherosclerosis, mitochondrial biogenesis is also modulated [62]. In human atherosclerotic arteries, the gene and protein PGC-1α levels are lower than in healthy subjects [63], especially in symptomatic plaques [64]. In human macrophages, PGC-1α overexpression in response to conjugated linoleic acid (CLA) treatment can block foam cell formation [64]. Vascular smooth muscle cell (VSMC) overexpression of PGC1α in rabbits diminished atherosclerotic lesions, including a decrease in proinflammatory cytokines, adhesion molecules, macrophage infiltration, senescence markers, matrix metalloproteinases (MMPs), and ROS production [65]. However, the role of PGC-1α in atherosclerosis is controversial. In APOE−/− mice, PGC-1α deficiency did not contribute to enhanced atherosclerosis [66]. In PGC-1α−/−ApoE−/− mice fed with a high-fat diet, the atheroprotective role of PGC-1α was limited to aged mice [67].
In the proteomic and genomic analysis on Fibulin-4R/R mouse aortas that resemble cutis laxa syndrome (associated with thoracic aortic aneurysm), the mitochondrial protein composition was altered and mitochondrial respiration reduced [68]. Similar results were observed in murine Loeys–Dietz syndrome [68]. Aortas from Fibulin-4R/R mice developed aneurysms and increased ROS production related to a decrease in PGC1α activity [68].
The modulation of PGC-1 coactivators after birth causes mitochondrial morphological derangement and progressive cardiomyopathy [69]. PGC-1α/β deficiency in the adult heart did not result in spontaneous abnormal mitochondrial dynamics or heart failure [69], but accelerated cardiac dysfunction and increased stress after pressure overload after transverse aortic constriction [70]. In contrast, the overexpression of PGC-1α in the vascular endothelial cells of mice induced re-endothelisalization after carotid injury [71]. Moreover, endothelial cells isolated from PGC-1α−/− mice showed an aberrant migration process [71].
PGC-1 coactivators may directly interact with transcription factor effectors such as the PPAR family, ERR family, Nuclear Respiratory Factor 1 (NRF-1), and mitochondrial transcription factor A (TFAM) [72]. PPAR family members have a critical role in mitochondrial biogenesis and FAO in the heart. PPARα deletion and overexpression regulate cardiomyocyte fatty acid metabolism and PPARα deficiency decreases ATP and promotes mitochondria abnormalities and cardiac fibrosis [73,74]. The estrogen-related receptors (ERRα, γ, etc.) are members of the orphan nuclear receptor (NR) family that governs aspects of mitochondrial function by regulating the expression of nuclear-encoded mitochondrial proteins [75]. ERRα knockout mice are susceptible to cardiac dysfunction in response to pressure overload and showed low ATP levels after ischemia [76]. By contrast, ERRγ knockout mice were not viable due to heart failure associated with abnormal oxidative metabolism [77]. NRF-1 and NRF-2 regulate the expression of the ETC complex [78] and activate the transcription of genes related to the replication and transcription of the mitochondrial genome [79]. NRF-1 or NRF-2 deficiency reduced mtDNA content and induced embryonic lethality in mice [80,81]. The cellular myelocytomatosis (c-Myc) transcription factor also controls mitochondrial function and induces NRF-1 and TFAM expression [82]. In mice under pressure overload after an ischemic insult, myocardial c-Myc activation modulated glucose metabolism and mitochondrial biogenesis, decreasing cardiac damage [83]. Mitochondrial biogenesis also modulates aneurysm formation. TFAM is a core mitochondrial transcription factor that has an essential role in mtDNA metabolism [84]. Patients with Marfan syndrome develop aortic aneurysms and have decreased TFAM expression and mtDNA levels in aortas [85]. In fact, in a transcriptomic analysis of aortas from Marfan syndrome (Fbn1c1039g/+) mice, the main pathway highlighted to be related to metabolic function was mitochondrial dysfunction. Moreover, loss of TFAM expression appears to be causative of aneurysm formation, as conditional TFAM-deficient VSMC mice lose their VSMC contractile capacity, develop aortic aneurysm formation, and die prematurely [85]. Restoration of mitochondrial metabolism with the NAD precursor nicotinamide riboside (NR) rapidly reduced aortic aneurysm formation in Marfan syndrome mice [85] and in the AngII-infused model in ApoE−/− mice fed with a western diet [86], increasing the expression of PGC-1α and TFAM. These results suggest a critical role of PGC-1α activity and its coactivators in CVD, however it could be controversial in some cases such as atherosclerotic disease (Figure 1).

4. Mitochondrial Dynamics

Mitochondria form a dynamic network in which mitochondria can modulate their morphology to create a tubular network co-ordinated by fission and fusion events to adapt their size, shape, and distribution to changing extracellular and intracellular environments. The balance between fission and fusion events is referred to as “mitochondrial dynamics” [87,88]. Numerous cellular activities, including the cell cycle, apoptotic or regulated necrosis cell death, metabolism, autophagy, and mitochondrial quality control, depend on rapid mitochondrial morphological modifications. The appropriate balance of these processes is essential for maintaining mitochondrial stability and function.
Mitochondrial fusion stimulates the assembly of individual mitochondria that combine their membranes. It takes two steps to fuse the outer and inner mitochondria membranes (IMM). Mitofusin 1 and 2 (MFN1 and MFN2) are located on the outside mitochondrial membrane (OMM) and regulate outer membrane fusion, and protein optic atrophy 1 (OPA1) regulates inner membrane fusion and cristae remodeling, an important determinant of mitochondrial metabolism [89,90].
Mitochondrial fission is the process of mitochondrial division into two separate mitochondrial organelles. Mitochondrial fission 1 (FIS1) and Dynamin-related protein 1 (DRP1) regulate mitochondrial fission. FIS1 is a single-pass transmembrane protein anchored to the outer mitochondrial membrane by its C-terminal region. It binds through MDV1 to DRP1, which is a member of the dynamin superfamily of GTPase-dynamin proteins. DRP1 is located in the cytosol which, during fission, is recruited to the mitochondria, driving membrane constriction in a GTP-dependent manner [91]. Other OMM proteins also participate in the recruitment of DRP1 as the mitochondrial fission factor (MFF), the mitochondrial dynamic proteins of 49 (MiD49) and 51 kDa (MiD51). These last two factors were enough to mediate the fission process in the absence of FIS1 and MFF [92]. Fission facilitates mitophagy, which is the breakdown and recycling of damaged mitochondria. DRP dysfunction may result in the impairment of the fission process, a situation that reduced mitophagy and, finally, induced the accumulation of mtDNA mutations and damaged proteins. In addition, fission-generated fragmented mitochondria produce higher amounts of reactive oxygen species, which may cause oxidative stress and cell injury [93].
Mitochondrial length is associated with the ability to perform OXPHOS. The blockade of mitochondrial fusion through MFN2-depletion boosts mitochondrial fragmentation, reduces mitochondrial membrane potential, lowers oxygen consumption and OXPHOS, and increases dependency on anaerobic glycolysis. Conversely, MFN2 over-expression stimulates mitochondrial metabolism and increases mitochondrial biogenesis and the expression of OXPHOS subunits [94,95]. OPA1 silencing results in similar effects, leading to mitochondrial fragmentation and decreased oxygen consumption via the decreased activity of mitochondrial complexes. At the molecular level, the GTP requirement of the effector proteins OPA1, MFN1, MFN2, and DRP1 may provide a direct link between the cell bioenergetic state and mitochondrial morphology, as most GTPase activities depend indirectly on the overall ATP content [96,97].

4.1. Role in Renal Damage

Mitochondrial injury has been linked to renal damage [98]. AKI is accompanied by excessive ROS production, promoting mitochondrial fission protein expression and activation (e.g., FIS1 and DRP1), mitochondrial fragmentation, renal tubular cell injury as well as death [12,98,99,100,101,102]. DRP1 activation and its mitochondrial translocation are an early feature of tubular cell injury, and the suppression of DRP-1 (dominant-negative mutant or siRNA) blocked mitochondrial fragmentation and apoptosis. Importantly, the DRP-1 inhibitor, mdivi-1, reduced mitochondrial fragmentation and protected the kidneys from both ischemia and cisplatin-induced AKI [103]. Mice from the conditional MFN2-deficient-kidney-epithelial-cell mice model have fewer nephrons, although kidney function was generally normal. Nevertheless, the primary proximal tubular MFS2-deficient cells displayed considerable mitochondrial fragmentation and were extremely vulnerable to apoptosis after ATP deprivation [104]. In murine AKI induced by ischemia-reperfusion, emodin was protective by regulating mitochondrial fission through the inhibition of DRP-1 activation [105]. Hence, boosting mitochondrial homeostasis can preserve kidney function. Cisplatin-induced nephrotoxicity is also associated with the increased DRP1-mediated mitochondrial fragmentation, cytochrome c release, and apoptosis of proximal tubular cells [106]. Indeed, the blockade of mitochondrial fission abrogated the cisplatin-induced proximal tubular cell apoptosis and improved kidney function, further supporting the critical role of mitochondrial dynamics in AKI [107].
TGFβ1 is a key driver of fibrosis in CKD. In cells with strong TGF1β1 expression, the mitochondrial fusion proteins OPA-1 and MFN2 are frequently downregulated, while the fission protein DRP1 is upregulated [108]. Furthermore, the DRP1 inhibitor mdivi-1 (50 mg/kg twice daily) was reported to reduce renal fibrosis induced by unilateral ureteral obstruction (UUO) for 10 days when administered either prophylactically or therapeutically (i.e., starting 3 days after UUO) [109]. However, despite the reported decreased fibrosis, kidney fibrosis was not quantified. Targeting DRP1 with a pharmacological inhibitor or siRNA increased mitochondrial respiration and reduced the fibroblast activation/proliferation and extracellular matrix production induced by TGFβ1 [109]. By contrast, in another report, Mdivi-1 enhanced the mtROS production induced by hypoxia-stimulated TGFβ1-Smad2/3 signaling in proximal tubular HK-2 cells and was reported to worsen renal fibrosis at day 7 after UUO when used at a dose of 2 mg/kg/day [110]. Overall, the impact of Mdivi-1 on UUO-induced kidney fibrosis remains controversial given the large difference in dose used by different authors and the fact that the report of decreased fibrosis is not supported by quantitative data. In diabetic kidney disease, mitochondrial fragmentation and FIS1 and DRP1 upregulation were reported to contribute to oxidative stress, cellular death, bioenergetic dysfunction, and poor mitochondrial autophagy and biogenesis [111]. FIS1 downregulation improved the mitochondrial morphology and reduced apoptosis in renal cells [111]. Moreover, DRP1 genetic deficiency or treatment with Mdvi1 or Berberine was nephroprotective in diabetic kidney disease [112,113,114]. These findings support the key role of renal mitochondria dynamics (fusion and fission) and associated proteins as therapeutic targets in CKD and AKI (Figure 2).

4.2. Role in Cardiovascular Disease

Balanced mitochondrial dynamics are particularly important for maintaining function in cardiovascular cells with high-energy demands, such as cardiomyocytes and VSMCs [115,116]. The role of mitochondrial dynamics during the cardiac development has been extensively studied in genetically modified mice. The systemic knockout of mitochondrial fusion proteins OPA1, MFN1, and MFN2 in mice are embryonically lethal [117,118], and the conditional combined deletion of MFN1/2 in adult cardiomyocytes triggers mitochondrial fragmentation, cardiac hypertrophy, and lethal dilated cardiomyopathy [119,120]. Moreover, gene-trapping MFN2 and OPA1 in mouse embryonic stem cells (ESCs) impaired cardiac development and inhibited cardiomyocyte differentiation [121]. Interestingly, mice with cardiac MFN1 deficiency maintained cardiac function and mitochondrial respiration [122], while cardiac MFN2 ablation resulted in dilated cardiomyopathy and hypertrophy [123,124]. The phenotypical differences between cardiac MFN1 and MFN2 deficiency may result from the role of MFN2 as a Parkin receptor, tethering the endoplasmic reticulum (ER) to mitochondria [125], which is essential for mitochondrial energy metabolism and calcium handling. Opa1delTTAG mutations in a cardiac model of ischemia-reperfusion injury is related to the elevated sensitivity to damage, imbalance in the dynamic mitochondrial Ca2+ uptake, and subsequent increase in the Na+/Ca2+ exchange (NCX) activity [126].
Mitochondrial fission is also critical for cardiac function. Cardiomyocyte-specific DRP1 knockout mice displayed lethal defects in mitochondrial respiration [127,128], likely dependent on overactivation of mitophagy [129] leading to dilated cardiomyopathy [130]. Moreover, mitochondrial fragmentation was observed in the myocardium of patients with heart failure and diabetic cardiomyopathy, in association with an increased expression of Bnip3, DRP1, and MFN1 [131,132].
In atherosclerosis, the expression of Mnf2 was reduced in thrombi from animal models and humans [133,134], while Mnf2 overexpression reduced atherosclerosis in rabbits [133,135]. Mitochondrial damage in VSMCs has been associated with plaque vulnerability in a model of atherogenesis accelerated by CKD [136]. In this sense, the proliferative VSMC phenotype has been associated with decreased MFN2 expression, while the inhibition of mitochondrial fission and of DRP1 reduced VSMC proliferation and migration in an ex vivo aortic ring assay and in rat carotid balloon injury, respectively [137,138]. Interestingly, endothelial dysfunction, a hallmark of atherosclerosis initiation, has also been associated with altered mitochondrial dynamics. Endothelial dysfunctional cells from a diabetic patient displayed the increased expression of Fis1 and fragmented mitochondria [139]. In addition, in a model of diabetic-induced atherosclerosis using ApoE knockout, inhibition of DRP1 reduced endothelial dysfunction, plaque formation [140], and smooth muscle cell calcification [141]. These results suggest that the inhibition of mitochondrial fission or stimulation of mitochondrial fusion may be of therapeutic interest in atherosclerosis.
Altered mitochondrial dynamics was also observed in cerebral ischemia. In a permanent middle cerebral artery occlusion (MCAO) model, MFN2 expression was reduced [142] and mitochondrial fission preceded neuronal loss [143]. Similarly, inhibition of DRP1 improved mitochondrial function, reduced mitochondrial fragmentation, and reduced infarct volume [144,145].
As is the case for mitochondrial biogenesis, there is evidence supporting the involvement of altered mitochondrial dynamics in aneurysm formation. In human abdominal aortic aneurysms, DRP1 expression is higher than in healthy donors [146,147]. Angiotensin II triggers mitochondrial fission in cultured VSMC [138], and this can be prevented by the DRP1 inhibitor Mdivi-1 [138]. In ApoE−/− knockout mice, Mdivi-1 attenuated the DRP1-mediated mitochondrial fission and inflammation in VSMCs and protected from abdominal aneurysms induced by AngII [146]. These results demonstrated that mitochondrial dynamics associated with fission and fusion re ais also critical in several pathological features of the cardiovascular disease (endothelial dysfunction, VSMCs calcification, cell hypertrophy, etc.) (Figure 2).

5. Oxidative Stress Originated in Mitochondria

ROS are produced by mitochondria during the normal functioning of the respiratory chain and OXPHOS, resulting in ATP production. ROS are molecular species with different levels of reactivity between them and need buffering by specific enzymatic systems as well as cellular macromolecules such as proteins and lipids that prevent their accumulation. Physiological ROS regulate pro-oxidative, anti-oxidative, or redox-sensitive signaling pathways including Mitogen-activated protein kinase (MAPK), phosphatidylinositol-3 kinase (PI3K), nuclear factor erythroid 2–related factor 2 (Nrf2), and Restriction Factor 1 (Ref1), but excessive ROS production or deficient scavenging results in oxidative stress and cell damage [148]. Conceptually, dysfunctional mitochondria cause mitochondrial oxidative stress (mtOS); however, oxidative stress may also stress the mitochondria [149] (Figure 3).

5.1. Role in Renal Damage

Oxidative stress has been involved in AKI and CKD by favoring a range of pathological mechanisms (cell death, inflammation, and fibrosis) following mitochondrial stress, and several antioxidant agents are kidney protective in preclinical AKI or CKD.
Preventive N-acetyl cysteine administration prevented decreased OXPHOS and FAO, mitochondrial depolarization, pro-oxidative stress, and a GSH deficit to protect mice from folic-acid-induced AKI [150,151]. The specific mitochondrial coenzyme Q10 antioxidants Mitoquinone mesylate (MitoQ) and Mitochondria-targeted carboxy-proxyl (Mito-CP) prevented cisplatin-induced AKI [152]. The natural antioxidant 2,3,5,6-tetramethylpyrazine (TMP) inhibited mitophagy, mitochondrial fragmentation, and oxidative stress and prevented contrast-induced AKI, decreasing tubular cell apoptosis, inflammation, and renal dysfunction [153]. Mitochondrial antioxidants also protected from AKI induced by sepsis, rhabdomyolysis, and ischemia-reperfusion [154,155,156]. The SkQ family of antioxidants, which have mitochondrial specificity and are used at a low dose, protected from ischemia-reperfusion AKI [157]. Mito-Tempo decreased mtROS in cultured tubular cells under oxidative stress and in vivo in ischemia-reperfusion AKI, preventing the decrease in TFAM levels and in the mtDNA copy number, decreasing mitochondrial fragmentation, restoring ATP production, and reducing cytokine release [158]. Mechanistically, TFAM deficiency in renal cells results in the mtDNA release into the cytosol and activation of the Sting pathway, a sentinel innate immune proinflammatory molecule that senses mislocated genomic or mtDNA. Sting promotes kidney damage following mtOS-dependent mtDNA oxidation [56,159,160].
TFAM deficiency is observed in aged murine kidneys [161]. Moreover, aged kidneys show Nrf2 inactivation and elevated lipid peroxidation and 8-OHdG levels resulting from oxidative stress in the context of dysfunctional mitochondria [157,162]. Innovative antioxidant therapies that specifically target mitochondria are being investigated as anti-aging compounds [163,164].
Elamipretide (SS-31) is a water-soluble peptide that accumulates in the inner mitochondrial membrane where it scavenges mtROS independently of mitochondrial membrane potential. SS-31 protects from acute ischemic or obstructive kidney damage. In murine diabetic nephropathy, elamipretide inhibited high-glucose-mediated ROS production, ATP depletion, ∆Ψm decay, and subsequent apoptotic mitochondrial pathway activation, decreasing the oxidative-stress-induced loss of glomerular cells (podocytes and endothelium), mesangial expansion, glomerulosclerosis, and release of renal proinflammatory and profibrotic cytokines [165,166,167]. Interestingly, mitochondrial targeting with Elamipretide protected against kidney ischemic injury in patients with atherosclerotic renal artery stenosis subjected to revascularization with percutaneous transluminal renal angioplasty, resulting in attenuated hypoxia, increased renal blood flow, and improved kidney function [168].
Clinical trials testing the mitochondrial antioxidants Elamipretide, Coenzyme Q10 (CoQ10), and MitoQ in renal diseases have been recently revised [35]. Of the 13 trials, six were completed, two were terminated, three are recruiting, and two have an unknown status [https://clinicaltrials.gov (accessed on 26 April 2023)]. Among them, one studied the role of Elamipretide in cardiac and renal damage in heart failure patients (NCT02914665). In the case of CoQ10, some trials in renal patients described its beneficial role in the oxidative response (NCT00908297), in the incidence of AKI (NCT04445779), and in the improvement of renal function (NCT04972552). MitoQ also has been tested as a dietary supplement for exercise capacity in patients with CKD and heart failure (NCT03960073) and for blood pressure reactivity and vascular function in healthy patients (NCT04334135).
Novel antioxidative nanoparticles (NPs) may also target mitochondria. Several organic (liposome-based or polymeric) and inorganic nanoparticles (QDs, gold nanoclusters (AuNCs), and CeO2) have physicochemical properties that may provide sustained and targeted drug delivery systems for dysfunctional mitochondria [169]. As an example, Ceria (CeO2) nanoparticles can scavenge ROS by reversible Ce+3-Ce+4 shuttling. Triphenylphosphonium-conjugated ceria nanoparticles localize into mitochondria and decreased mitochondrial damage in Alzheimer’s disease [170]. Moreover, NPs carrying several compounds such as resveratrol, quercetin, and melatonin scavenge ROS during kidney injury [171,172,173]. These finding demonstrated that the modulation of the mitochondrial oxidative response with several antioxidant compounds, vehicularized or not by NP, have beneficial effects in inflammation, cell death, and fibrosis in renal damage.

5.2. Role in Cardiovascular Damage

Mitochondrial DNA damage and oxidative stress have been observed in CVDs [174,175,176,177]. OXPHOS and ATP generation are the main processes that produce ROS in the heart [178]. The blockade of ETC complex I by site IQ electron leak (S1QELs) suppressors inhibit ROS production, protected cardiac function, and decreased infarct size following ischemia-reperfusion [179].
mtROS have a key role in cardiovascular aging. In mice, catalase overexpression (mCAT) in mitochondria reduced cardiac aging [180]. In addition, mice with mCAT overexpression are resistant to angiotensin-II-induced cardiac hypertrophy, fibrosis, and heart failure [181]. In murine metabolic heart disease caused by a high-fat, high-sucrose (HFHS) diet, the transgenic overexpression of catalase in mitochondria prevented contractile dysfunction in the heart [182]. In fact, ROS can decrease the mtDNA copy number by damaging mtDNA replication enzymes [183]. The mtDNA copy number is inversely related to the prevalence/incidence of cardiovascular disease and sudden cardiac death [184,185].
In atherosclerosis, increased ROS production leads to DNA, protein, and lipid oxidation, mainly located in plaque VSMCs [186,187]. mtDNA damage precedes and correlates with atherosclerotic plaque development [188,189]. However, there are controversial opinions about the role of mtROS in mtDNA damage in the context of atherosclerosis progression. Mice with mitochondrial mutations showed no increase in ROS generation despite extensive mtDNA defects [190,191]. In addition, the antioxidant compound MitoQ reduced metabolic syndrome features but had no impact on murine atherosclerosis development [192]. More studies are needed to clarify the role of mtROS in atherosclerosis plaque formation and progression.
A role of mtROS in cardiac disease is better supported by the evidence [175]. In a Chinese population, there was an inverse association between the mtDNA copy number, coronary artery disease, and ROS production [193]. A genomic TFAM blockade caused mitochondrial dysfunction, increased ROS production, and decreased cardiomyocyte proliferation, which were rescued by the mtROS scavenger MitoTEMPO [194]. In cardiac ischemia-reperfusion injury, ROS oxidized thiols in mitochondrial respiratory complex I, inducing an excess ROS production during reperfusion [195]. Mitochondria-specific antioxidant compounds such as 10-(6-Plastoquinonyl) decyltriphenyl-phosphonium (SkQ), Elamipretide, the α-lipoic acid analogue CMX-2043, and MitoQ were protective in cardiac ischemia/reperfusion injury [196,197,198,199,200]. MitoQ also protected endothelial cells from damage induced by chronic exposure to nitroglycerine [201] and controlled blood pressure in hypertensive rats [202]. The SOD/catalase mimetic antioxidant Euk-8 decreased oxidative stress and pressure-overload-induced heart failure in harlequin (Hq) mutant mice [203].
In addition, several lines of evidence indicate that excessive oxidative damage induced by mtROS occurs in atherosclerotic lesions of both animal models and humans [204]. Moreover, mitochondria-targeted anti-oxidants are growing as a new class of compounds for age-related diseases, including atherosclerosis. In this sense, mitochondrial-targeted antioxidant MitoQ ameliorates inflammation, radical oxygen species (ROS) production, and leukocyte–endothelium interactions in atherosclerosis of Type 2 Diabetes (T2D) patients [205]. In addition, mitochondria-targeted esculetin (mito-Esc) antioxidant diminish atherosclerosis in aged ApoE−/− by delaying vascular senescence and pro-inflammatory processes, and by improving mitochondrial biogenesis and function [206]. All these results showed the key role of the mitochondrial oxidative response in the modulation of pathological features of cardiovascular damage including vascular senescence, mtDNA damage, as well as pro-inflammatory processes.

6. Mitophagy

Mitophagy promotes mitochondria recycling through the engulfment of whole or fragmented mitochondria by autophagosomes, transferring the mitochondrial cargo to lysosomes for degradation [207,208,209].
Most studies investigating the role of mitophagy in physiological or pathological conditions focus on the PTEN Induced Kinase 1 (PINK1)-Parkin RBR E3 Ubiquitin Protein Ligase (Parkin) pathway, which is the most classic mitophagy pathway. PINK1 is a serine kinase that can accumulate in the outer mitochondrial membrane. Upon the loss of mitochondrial membrane potential (ΔΨm), cytosolic Parkin is recruited to the mitochondria by PINK1, resulting in the ubiquitination of mitochondrial proteins, aggregation of p62, and specific binding of microtubule-associated protein 1A/1B-light chain 3 (LC3), thus promoting the recruitment of the autophagosome membrane [210] and initiating the clearance of damaged mitochondria [211]. Mitophagy may also be activated in a PINK-independent manner by mitophagy receptors located in the outer mitochondrial membrane that trigger mitophagy through direct interactions with LC3 on the autophagosome [212]. Nix, BCL2/adenovirus E1B 19 kDa protein-interacting protein 3 (Bnip3), FUN14 domain containing 1 (Fundc1), and others function as mitophagy receptors [213,214,215,216,217,218] (Figure 4).

6.1. Role in Renal Damage

In CKD patients and experimental kidney disease, mitophagy-related genes such as PINK1 and Parkin are downregulated [219], and Pink1 and Parkin deletion blocks mitophagy, alters mitochondrial homeostasis, and increases renal fibrosis [220]. In adenine-induced CKD, the myeloid-cell-specific deletion of MFN2 decreased mitophagy, magnified macrophage-derived profibrotic responses, and impaired renal function [220]. In cultured HK-2 cells, Parkin siRNA increased albumin-induced mitochondrial dysfunction and oxidative stress, which were decreased by Parkin overexpression [221]. The single or double Pink1- and Parkin-KO mice display an increased sensitivity to mitochondrial damage, inflammation, and ROS production following kidney ischemia-reperfusion [222]. Murine cisplatin-induced AKI was also more severe in PINK1- and Parkin-deficient mice [223]. Tubular epithelial cells deficient in Pink1/Parkin were more sensitive to cisplatin-induced mitochondrial dysfunction and apoptosis [224]. The deficiency of PINK1/Parkin2 also triggered mitochondrial damage and inflammatory responses [222].
In murine kidney ischemia-reperfusion injury, BNIP3 overexpression induced mitophagy and protected from tubular renal damage [225]. Autophagy-related-5 (ATG5) and autophagy-related-7 (ATG7) are autophagy-related proteins with a key role in autophagic vesicle formation, reducing the degraded macromolecules and organelles which accumulated in the tissue after damage [226]. ATG5 or ATG7 mutant mice displayed deficient mitophagy and developed mitochondrial dysfunction and podocyte and tubular cell vacuolization leading to focal segmental glomerulosclerosis and organ failure [227]. In rats, high-calorie diets decreased mitophagy, caused an abnormal mitochondrial morphology, and increased age-associated kidney injury, while calorie restriction was protective [228].
Altered mitophagy may also contribute to DKD. Optineurin (OPTN) and PINK1 levels were decreased in human diabetic patients; the Mdivi-1 increased renal senescence, while Torin1, an autophagy/mitophagy agonist, was protective [229]. In diabetic mice, Src activation increased the severity of podocyte injury via the suppression of FUNDC1-mediated mitophagy [230]. The antioxidant drug MitoQ upregulated PINK and Parkin expression, preventing tubular injury in diabetic kidney disease (DKD) in vitro and in vivo [231]. The anthelmintic drug niclosamide and its analogues, which are indirect PINK1 activators, exerted renoprotective effects in preclinical type 2 (db/db) and type 1 (streptozotocin) diabetes and also in kidney injury induced by ischemia-reperfusion, UUO, and Adriamycin damage [232,233,234,235,236]. In cultured podocytes, silencing the PINK1/Parkin mitophagy pathway dramatically aggravated mitochondrial ROS and palmitic-acid-induced apoptosis [237]. A bioactive constituent of Fenugreek (Orientin) restored mitochondrial autophagy and reduced high-glucose-induced podocyte apoptosis [238]. Finally, the upregulation of mitophagy components such as PINK1, PARKIN, and related proteins (ATG5 or ATG7) by pharmacological approaches (autophagy agonists, anthelmintic drugs, antioxidants, etc.) restored the damage induced by AKI and CKD (including glomerular damage).

6.2. Role in Cardiovascular Damage

Several studies have related autophagic or mitophagy defects with the development of specific cardiovascular disorders [208]. Proteins associated with autophagy, including Beclin1 and different ATG components, were modulated in experimental cardiovascular damage [239]. Heterocygote Becn1+/− mice were more resistant to cardiac damage induced by myocardial infarction than controls [240]. ATG7-overexpressing mice had milder ventricular dysfunction, and cardiac fibrosis and hypertrophy in desmin-related cardiomyopathy [241]. In the hypertensive model of Angiotensin II infusion, Atg5+/− mice exhibited a dramatic cardiac hypertrophy compared to AngII-infused mice [242]. In contrast, the overexpression of other autophagy components could be detrimental. Overexpression of RHEB (a GTP-binding protein that inhibits autophagy by activating mTORC1) in cardiomyocytes increased cardiac injury following myocardial infarction [243].
The PINK1 (PTEN-induced kinase 1)–Parkin (also known as PARK2) signaling pathway is a key receptor-independent pathway that controls mitophagy [244]. PINK1 in mitochondria undergoes rapid proteolytic degradation under normal conditions [72]. In murine cardiac ischemia-reperfusion injury, Pgam5 (PGAM family member 5, mitochondrial serine/threonine protein phosphatase) deletion increased the infarct size [245]. Pgam5 participates in mitophagy. In Park2−/− mice, simvastatin did not activate mitophagy nor protect from ischemia/reperfusion injury [246]. Park2/ mice also showed reduced cardiomyocyte mitophagy and more severe cardiac damage following myocardial infarction [247].
Some protein receptors mediate mitophagy independent of the PINK1–Parkin signaling pathway. Deficiency of the autophagy protein receptor B-cell leukemia/lymphoma 2 protein (BCL2) interacting protein 3 (Bnip3) in mice reduced myocardial damage and preserved the heart structure following cardiac ischemia/reperfusion injury [248]. The deletion of FUNDC1 in mice protected from cardiac ischemia/reperfusion injury through the regulation of platelet activity [249].
In addition, pharmacological inducers of autophagy/mitophagy, such as rapamycin, protected from experimental myocardial infarction [250]. In contrast, autophagy inhibitors as 3–MA and BafA1 exacerbated postinfarction cardiac remodeling and dysfunction [250,251].
There is also a link between mitochondrial dysfunction and atherosclerosis associated with oxidative stress, apoptosis, and deficient ATP synthesis [57]. In atherosclerotic mice (Apoe−/−), the endothelial-specific deletion of Atg5 increased endothelial cell apoptosis and atherosclerotic lesions [252]. The overexpression of nuclear receptor subfamily 4 group A member 1 (NR4A1) increased the severity of atherosclerosis through the mitophagy exacerbation that diminished the number of mitochondria and increased endothelial apoptosis [253]. PINK1−/− mice were protected from atherosclerosis and had reduced mitophagy [254]. By contrast, the overexpression of PINK1 or Parkin restored oxLDL-induced VSMC apoptosis in atherosclerosis [57]. The atherosclerotic microenvironment activated rapamycin signaling in macrophages, blocking mitophagy [255]. Other mitophagy protein such as NIX (Bnip3L) that directly interacts with LC3 regulated platelet mitophagy. Nix−/− mice displayed less platelet activation and aggregation and a longer time to complete the thrombotic occlusion of the carotid artery induced by FeCl3 [256]. In atherosclerotic aged mice, the mitophagy inducer spermidine reduced mitochondrial dysfunction and atherosclerosis [257]. The blockade of mitophagy induces cardiovascular damage and the overexpression of mitophagy markers improves this injury. However, autophagy upregulation currently is controversial because an excessive activation of this degradative process could have detrimental effects in vascular tissue.

7. Conclusions

Mitochondria are dynamic and heterogeneous organelles that are the most important source of oxidative stress and ATP production, and regulate cell function, survival, and death. Mitochondrial dysfunction is involved in the pathophysiology of CKD and CVD. However, a limitation of this review is the paucity of studies that shed light on how mitochondrial dysfunction contributes to the development and progression of the main vascular pathologies that could be affected by kidney injury and vice versa.
Mitochondrial structural/functional abnormalities such as impaired mitochondrial biogenesis, defective mitochondrial dynamics, mitochondrial dysfunction, and oxidative stress contribute to the development and progression of CRS. Several therapeutic approaches using pharmacological strategies to preserve mitochondrial processes successfully slowed the progression of cardiovascular and renal damage in preclinical models. An improved understanding of the molecular links between mitochondrial dysfunction and CVD–CKD interplay may pave the way for clinical studies that test therapeutic interventions aimed at preserving mitochondrial function to protect from CRS.

Author Contributions

Conceptualization, S.R.-M.; N.M.-B.; writing—original draft preparation, S.R.-M.; N.M.-B.; J.O.; A.B.S.; A.O.; A.M.R. and M.R.-O.; writing—review and editing S.R.-M.; N.M.-B.; J.O.; A.B.S.; A.O.; A.M.R. and M.R.-O.; visualization, S.R.-M.; N.M.-B.; J.O.; A.B.S.; A.O.; A.M.R. and M.R.-O.; project administration, S.R.-M.; funding acquisition S.R.-M.; N.M.-B.; A.B.S.; A.O.; A.M.R. and M.R.-O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by grants from the Instituto de Salud Carlos III (ISCIII) and Fondos FEDER European Union (PI20/00140, PI19/00815, and DTS20/00083, PI19/00588, PI19/00815, PI22/00469, PI22/00050, PI18/01133, PI21/01453, PI21/01126, Red de Investigación Renal REDINREN: RD16/0009/0003, and RICORS2040; RD21/0005/0002, RD21/0005/0001) funded by European Union—NextGenerationEU, Sociedad Española de Nefrología, Sociedad española de aterosclerosis (FEA 2021-BIB 05-21), and 2021 PREMIOS A LA INVESTIGACIÓN L’ORÉAL-UNESCO For Women in Science. Innovation programme under the Marie Skłodowska-Curie grant of the European Union’s Horizon 2020 (IMProve-PD ID: 812699) to M.R.-O., and the Juan de la Cierva incorporacion grant: IJC2018-035187-I to S.R.-M. J.O. is supported by a Ramón y Cajal contract (RYC2021-033343-I), from the Spanish Ministry of Science and Innovation and granted by Fundación MERCK—“Fundación FEDER de investigación clínica en enfermedades raras 2022” and by “V-Ayudas Muévete por los que no pueden 2021”. A.B.S was supported by the MICINN Ramon y Cajal program. The ISCIII Miguel Servet program supported N.M.-B.

Institutional Review Board Statement

Not applicable.

Acknowledgments

We thank Laura Almenara for her administrative support to our research group.

Conflicts of Interest

A.O. has received grants from Sanofi and consultancy, speaker fees, or travel support from Advicciene, Astellas, AstraZeneca, Amicus, Amgen, Fresenius Medical Care, GSK, Bayer, Sanofi-Genzyme, Menarini, Kyowa Kirin, Alexion, Idorsia, Chiesi, Otsuka, Novo-Nordisk, and Vifor Fresenius Medical Care Renal Pharma, and is Director of the Catedra Mundipharma-UAM for diabetic kidney disease and the Catedra Astrazeneca-UAM for chronic kidney disease and electrolytes. He also has stock in Telara Farma. The other authors declare no competing interest.

References

  1. Ronco, C.; Bellasi, A.; Di Lullo, L. Cardiorenal Syndrome: An Overview. Adv. Chronic Kidney Dis. 2018, 25, 382–390. [Google Scholar] [CrossRef]
  2. Ronco, C.; McCullough, P.; Anker, S.D.; Anand, I.; Aspromonte, N.; Bagshaw, S.M.; Bellomo, R.; Berl, T.; Bobek, I.; Cruz, D.N.; et al. Cardio-renal syndromes: Report from the consensus conference of the acute dialysis quality initiative. Eur. Heart J. 2010, 31, 703–711. [Google Scholar] [CrossRef] [PubMed]
  3. Melenovsky, V.; Cervenka, L.; Viklicky, O.; Franekova, J.; Havlenova, T.; Behounek, M.; Chmel, M.; Petrak, J. Kidney Response to Heart Failure: Proteomic Analysis of Cardiorenal Syndrome. Kidney Blood Press. Res. 2018, 43, 1437–1450. [Google Scholar] [CrossRef] [PubMed]
  4. Hill, N.R.; Fatoba, S.T.; Oke, J.L.; Hirst, J.A.; O’Callaghan, C.A.; Lasserson, D.S.; Hobbs, F.D.R. Global Prevalence of Chronic Kidney Disease—A Systematic Review and Meta-Analysis. PLoS ONE 2016, 11, e0158765. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Jager, K.J.; Kovesdy, C.; Langham, R.; Rosenberg, M.; Jha, V.; Zoccali, C. A single number for advocacy and communication-worldwide more than 850 million individuals have kidney diseases. Kidney Int. 2019, 96, 1048–1050. [Google Scholar] [CrossRef]
  6. Wheeler, D.C.; Stefánsson, B.V.; Jongs, N.; Chertow, G.M.; Greene, T.; Hou, F.F.; McMurray, J.J.V.; Correa-Rotter, R.; Rossing, P.; Toto, R.D.; et al. Effects of dapagliflozin on major adverse kidney and cardiovascular events in patients with diabetic and non-diabetic chronic kidney disease: A prespecified analysis from the DAPA-CKD trial. Lancet Diabetes Endocrinol. 2021, 9, 22–31. [Google Scholar] [CrossRef]
  7. Thompson, S.; James, M.; Wiebe, N.; Hemmelgarn, B.; Manns, B.; Klarenbach, S.; Tonelli, M. Cause of Death in Patients with Reduced Kidney Function. J. Am. Soc. Nephrol. 2015, 26, 2504–2511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Lekawanvijit, S.; Krum, H. Cardiorenal syndrome: Acute kidney injury secondary to cardiovascular disease and role of protein-bound uraemic toxins. J. Physiol. 2014, 592, 3969–3983. [Google Scholar] [CrossRef]
  9. Pliquett, R.U. Cardiorenal Syndrome: An Updated Classification Based on Clinical Hallmarks. J. Clin. Med. 2022, 11, 2896. [Google Scholar] [CrossRef]
  10. Doenst, T.; Nguyen, T.D.; Abel, E.D. Cardiac metabolism in heart failure: Implications beyond ATP production. Circ. Res. 2013, 113, 709–724. [Google Scholar] [CrossRef] [Green Version]
  11. Duann, P.; Lin, P.H. Mitochondria Damage and Kidney Disease. Adv. Exp. Med. Biol. 2017, 982, 529–551. [Google Scholar] [CrossRef] [PubMed]
  12. Brooks, C.; Wei, Q.; Cho, S.G.; Dong, Z. Regulation of mitochondrial dynamics in acute kidney injury in cell culture and rodent models. J. Clin. Investig. 2009, 119, 1275–1285. [Google Scholar] [CrossRef] [PubMed]
  13. Howell, G.M.; Gomez, H.; Collage, R.D.; Loughran, P.; Zhang, X.; Escobar, D.A.; Billiar, T.R.; Zuckerbraun, B.S.; Rosengart, M.R. Augmenting autophagy to treat acute kidney injury during endotoxemia in mice. PLoS ONE 2013, 8, e69520. [Google Scholar] [CrossRef] [PubMed]
  14. Clementi, A.; Virzì, G.M.; Brocca, A.; De Cal, M.; Pastori, S.; Clementi, M.; Granata, A.; Vescovo, G.; Ronco, C. Advances in the pathogenesis of cardiorenal syndrome type 3. Oxid. Med. Cell. Longev. 2015, 2015, 148082. [Google Scholar] [CrossRef] [Green Version]
  15. Di Lullo, L.; Bellasi, A.; Russo, D.; Cozzolino, M.; Ronco, C. Cardiorenal acute kidney injury: Epidemiology, presentation, causes, pathophysiology and treatment. Int. J. Cardiol. 2017, 227, 143–150. [Google Scholar] [CrossRef]
  16. Shi, S.; Zhang, B.; Li, Y.; Xu, X.; Lv, J.; Jia, Q.; Chai, R.; Xue, W.; Li, Y.; Wang, Y.; et al. Mitochondrial Dysfunction: An Emerging Link in the Pathophysiology of Cardiorenal Syndrome. Front. Cardiovasc. Med. 2022, 9, 837270. [Google Scholar] [CrossRef]
  17. Popkov, V.A.; Silachev, D.N.; Zalevsky, A.O.; Zorov, D.B.; Plotnikov, E.Y. Mitochondria as a Source and a Target for Uremic Toxins. Int. J. Mol. Sci. 2019, 20, 3094. [Google Scholar] [CrossRef] [Green Version]
  18. Zhang, A.; Jia, Z.; Wang, N.; Tidwell, T.J.; Yang, T. Relative contributions of mitochondria and NADPH oxidase to deoxycorticosterone acetate-salt hypertension in mice. Kidney Int. 2011, 80, 51–60. [Google Scholar] [CrossRef] [Green Version]
  19. Ravarotto, V.; Bertoldi, G.; Innico, G.; Gobbi, L.; Calò, L.A. The Pivotal Role of Oxidative Stress in the Pathophysiology of Cardiovascular-Renal Remodeling in Kidney Disease. Antioxidants 2021, 10, 1041. [Google Scholar] [CrossRef]
  20. Morciano, G.; Boncompagni, C.; Ramaccini, D.; Pedriali, G.; Bouhamida, E.; Tremoli, E.; Giorgi, C.; Pinton, P. Comprehensive Analysis of Mitochondrial Dynamics Alterations in Heart Diseases. Int. J. Mol. Sci. 2023, 24, 3414. [Google Scholar] [CrossRef]
  21. Ikeda, Y.; Shirakabe, A.; Brady, C.; Zablocki, D.; Ohishi, M.; Sadoshima, J. Molecular mechanisms mediating mitochondrial dynamics and mitophagy and their functional roles in the cardiovascular system. J. Mol. Cell. Cardiol. 2015, 78, 116–122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Bhargava, P.; Schnellmann, R.G. Mitochondrial energetics in the kidney. Nat. Rev. Nephrol. 2017, 13, 629–646. [Google Scholar] [CrossRef] [PubMed]
  23. Zhu, Z.; Hu, J.; Chen, Z.; Feng, J.; Yang, X.; Liang, W.; Ding, G. Transition of acute kidney injury to chronic kidney disease: Role of metabolic reprogramming. Metabolism 2022, 131, 155194. [Google Scholar] [CrossRef] [PubMed]
  24. Mallick, R.; Duttaroy, A.K. Modulation of endothelium function by fatty acids. Mol. Cell. Biochem. 2022, 477, 15–38. [Google Scholar] [CrossRef]
  25. Palm, C.L.; Nijholt, K.T.; Bakker, B.M.; Westenbrink, B.D. Short-Chain Fatty Acids in the Metabolism of Heart Failure—Rethinking the Fat Stigma. Front. Cardiovasc. Med. 2022, 9, 915102. [Google Scholar] [CrossRef] [PubMed]
  26. Su, L.; Zhang, J.; Gomez, H.; Kellum, J.A.; Peng, Z. Mitochondria ROS and mitophagy in acute kidney injury. Autophagy 2023, 19, 2084862. [Google Scholar] [CrossRef] [PubMed]
  27. Liao, X.; Huang, C.; Zhang, D.; Wang, J.; Li, J.; Jin, H.; Huang, C. Mitochondrial catalase induces cells transformation through nucleolin-dependent Cox-2 mRNA stabilization. Free Radic. Biol. Med. 2017, 113, 478–486. [Google Scholar] [CrossRef]
  28. Suski, J.M.; Lebiedzinska, M.; Bonora, M.; Pinton, P.; Duszynski, J.; Wieckowski, M.R. Relation between mitochondrial membrane potential and ROS formation. Methods Mol. Biol. 2012, 810, 183–205. [Google Scholar] [CrossRef] [PubMed]
  29. Birk, A.V.; Chao, W.M.; Bracken, C.; Warren, J.D.; Szeto, H.H. Targeting mitochondrial cardiolipin and the cytochrome c/cardiolipin complex to promote electron transport and optimize mitochondrial ATP synthesis. Br. J. Pharmacol. 2014, 171, 2017–2028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Wan, J.; Kalpage, H.A.; Vaishnav, A.; Liu, J.; Lee, I.; Mahapatra, G.; Turner, A.A.; Zurek, M.P.; Ji, Q.; Moraes, C.T.; et al. Regulation of Respiration and Apoptosis by Cytochrome c Threonine 58 Phosphorylation. Sci. Rep. 2019, 9, 15815. [Google Scholar] [CrossRef] [Green Version]
  31. Tang, C.; Cai, J.; Yin, X.M.; Weinberg, J.M.; Venkatachalam, M.A.; Dong, Z. Mitochondrial quality control in kidney injury and repair. Nat. Rev. Nephrol. 2021, 17, 299–318. [Google Scholar] [CrossRef]
  32. Jornayvaz, F.R.; Shulman, G.I. Regulation of mitochondrial biogenesis. Essays Biochem. 2010, 47, 69–84. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Ploumi, C.; Daskalaki, I.; Tavernarakis, N. Mitochondrial biogenesis and clearance: A balancing act. FEBS J. 2017, 284, 183–195. [Google Scholar] [CrossRef] [PubMed]
  34. Fontecha-barriuso, M.; Martin-sanchez, D.; Martinez-moreno, J.M.; Monsalve, M.; Ramos, A.M.; Sanchez-niño, M.D.; Ruiz-ortega, M.; Ortiz, A.; Sanz, A.B. The Role of PGC-1α and Mitochondrial Biogenesis in Kidney Diseases. Biomolecules 2020, 10, 347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Fontecha-Barriuso, M.; Lopez-Diaz, A.M.; Guerrero-Mauvecin, J.; Miguel, V.; Ramos, A.M.; Sanchez-Niño, M.D.; Ruiz-Ortega, M.; Ortiz, A.; Sanz, A.B. Tubular Mitochondrial Dysfunction, Oxidative Stress, and Progression of Chronic Kidney Disease. Antioxidants 2022, 11, 1356. [Google Scholar] [CrossRef] [PubMed]
  36. Riehle, C.; Abel, E.D. PGC-1 proteins and heart failure. Trends Cardiovasc. Med. 2012, 22, 98–105. [Google Scholar] [CrossRef] [Green Version]
  37. Patten, I.S.; Arany, Z. PGC-1 coactivators in the cardiovascular system. Trends Endocrinol. Metab. 2012, 23, 90–97. [Google Scholar] [CrossRef]
  38. Finck, B.N.; Kelly, D.P. PGC-1 coactivators: Inducible regulators of energy metabolism in health and disease. J. Clin. Investig. 2006, 116, 615–622. [Google Scholar] [CrossRef] [Green Version]
  39. Transcriptional Control of Mitochondrial Energy Metabolism through the PGC1 Coactivators—PubMed. Available online: https://pubmed.ncbi.nlm.nih.gov/18074631/ (accessed on 22 April 2023).
  40. Ruiz-Andres, O.; Suarez-Alvarez, B.; Sánchez-Ramos, C.; Monsalve, M.; Sanchez-Niño, M.D.; Ruiz-Ortega, M.; Egido, J.; Ortiz, A.; Sanz, A.B. The inflammatory cytokine TWEAK decreases PGC-1α expression and mitochondrial function in acute kidney injury. Kidney Int. 2016, 89, 399–410. [Google Scholar] [CrossRef] [Green Version]
  41. Tran, M.; Tam, D.; Bardia, A.; Bhasin, M.; Rowe, G.C.; Kher, A.; Zsengeller, Z.K.; Reza Akhavan-Sharif, M.; Khankin, E.V.; Saintgeniez, M.; et al. PGC-1α promotes recovery after acute kidney injury during systemic inflammation in mice. J. Clin. Investig. 2011, 121, 4003–4014. [Google Scholar] [CrossRef] [Green Version]
  42. Portilla, D.; Dai, G.; McClure, T.; Bates, L.; Kurten, R.; Megyesi, J.; Price, P.; Li, S. Alterations of PPARalpha and its coactivator PGC-1 in cisplatin-induced acute renal failure. Kidney Int. 2002, 62, 1208–1218. [Google Scholar] [CrossRef] [PubMed]
  43. Oh, C.J.; Ha, C.M.; Choi, Y.K.; Park, S.; Choe, M.S.; Jeoung, N.H.; Huh, Y.H.; Kim, H.J.; Kweon, H.S.; Lee, J.; et al. Pyruvate dehydrogenase kinase 4 deficiency attenuates cisplatin-induced acute kidney injury. Kidney Int. 2017, 91, 880–895. [Google Scholar] [CrossRef]
  44. Fontecha-Barriuso, M.; Martín-Sánchez, D.; Martinez-Moreno, J.M.; Carrasco, S.; Ruiz-Andrés, O.; Monsalve, M.; Sanchez-Ramos, C.; Gómez, M.J.; Ruiz-Ortega, M.; Sánchez-Niño, M.D.; et al. PGC-1α deficiency causes spontaneous kidney inflammation and increases the severity of nephrotoxic AKI. J. Pathol. 2019, 249, 65–78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Tran, M.T.; Zsengeller, Z.K.; Berg, A.H.; Khankin, E.V.; Bhasin, M.K.; Kim, W.; Clish, C.B.; Stillman, I.E.; Karumanchi, S.A.; Rhee, E.P.; et al. PGC1α drives NAD biosynthesis linking oxidative metabolism to renal protection. Nature 2016, 531, 528–532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Campbell, S.; Michaelson, J.; Burkly, L.; Putterman, C. The role of TWEAK/Fn14 in the pathogenesis of inflammation and systemic autoimmunity. Front. Biosci. 2004, 9, 2273–2284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Wiley, S.R.; Winkles, J.A. TWEAK, a member of the TNF superfamily, is a multifunctional cytokine that binds the TweakR/Fn14 receptor. Cytokine Growth Factor Rev. 2003, 14, 241–249. [Google Scholar] [CrossRef] [PubMed]
  48. Sanz, A.B.; Sanchez-Nĩo, M.D.; Ortiz, A. TWEAK, a multifunctional cytokine in kidney injury. Kidney Int. 2011, 80, 708–718. [Google Scholar] [CrossRef] [Green Version]
  49. Ortiz, A.; Sanz, A.B.; García, B.M.; Moreno, J.A.; Niño, M.D.S.; Martín-Ventura, J.L.; Egido, J.; Blanco-Colio, L.M. Considering TWEAK as a target for therapy in renal and vascular injury. Cytokine Growth Factor Rev. 2009, 20, 251–258. [Google Scholar] [CrossRef]
  50. Sanz, A.B.; Justo, P.; Sanchez-Nino, M.D.; Blanco-Colio, L.M.; Winkles, J.A.; Kreztler, M.; Jakubowski, A.; Blanco, J.; Egido, J.; Ruiz-Ortega, M.; et al. The cytokine TWEAK modulates renal tubulointerstitial inflammation. J. Am. Soc. Nephrol. 2008, 19, 695–703. [Google Scholar] [CrossRef] [Green Version]
  51. Sharma, K.; Karl, B.; Mathew, A.V.; Gangoiti, J.A.; Wassel, C.L.; Saito, R.; Pu, M.; Sharma, S.; You, Y.H.; Wang, L.; et al. Metabolomics reveals signature of mitochondrial dysfunction in diabetic kidney disease. J. Am. Soc. Nephrol. 2013, 24, 1901–1912. [Google Scholar] [CrossRef] [Green Version]
  52. Kang, H.M.; Ahn, S.H.; Choi, P.; Ko, Y.A.; Han, S.H.; Chinga, F.; Park, A.S.D.; Tao, J.; Sharma, K.; Pullman, J.; et al. Defective fatty acid oxidation in renal tubular epithelial cells has a key role in kidney fibrosis development. Nat. Med. 2015, 21, 37–46. [Google Scholar] [CrossRef] [PubMed]
  53. Han, S.H.; Wu, M.Y.; Nam, B.Y.; Park, J.T.; Yoo, T.H.; Kang, S.W.; Park, J.; Chinga, F.; Li, S.Y.; Susztak, K. PGC-1 α Protects from Notch-Induced Kidney Fibrosis Development. J. Am. Soc. Nephrol. 2017, 28, 3312–3322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Li, S.Y.; Park, J.; Qiu, C.; Han, S.H.; Palmer, M.B.; Arany, Z.; Susztak, K. Increasing the level of peroxisome proliferator-activated receptor γ coactivator-1α in podocytes results in collapsing glomerulopathy. JCI Insight 2017, 2, 92930. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Koh, E.; Kim, Y.K.; Shin, D.; Kim, K.S. MPC1 is essential for PGC-1α-induced mitochondrial respiration and biogenesis. Biochem. J. 2018, 475, 1687–1699. [Google Scholar] [CrossRef]
  56. Chung, K.W.; Dhillon, P.; Huang, S.; Sheng, X.; Shrestha, R.; Qiu, C.; Kaufman, B.A.; Park, J.; Pei, L.; Baur, J.; et al. Mitochondrial Damage and Activation of the STING Pathway Lead to Renal Inflammation and Fibrosis. Cell Metab. 2019, 30, 784–799.e5. [Google Scholar] [CrossRef] [PubMed]
  57. Chistiakov, D.A.; Shkurat, T.P.; Melnichenko, A.A.; Grechko, A.V.; Orekhov, A.N. The role of mitochondrial dysfunction in cardiovascular disease: A brief review. Ann. Med. 2018, 50, 121–127. [Google Scholar] [CrossRef]
  58. Friedman, J.R.; Nunnari, J. Mitochondrial form and function. Nature 2014, 505, 335–343. [Google Scholar] [CrossRef] [Green Version]
  59. Andersson, U.; Scarpulla, R.C. Pgc-1-related coactivator, a novel, serum-inducible coactivator of nuclear respiratory factor 1-dependent transcription in mammalian cells. Mol. Cell. Biol. 2001, 21, 3738–3749. [Google Scholar] [CrossRef] [Green Version]
  60. Lehman, J.J.; Barger, P.M.; Kovacs, A.; Saffitz, J.E.; Medeiros, D.M.; Kelly, D.P. Peroxisome proliferator-activated receptor gamma coactivator-1 promotes cardiac mitochondrial biogenesis. J. Clin. Investig. 2000, 106, 847–856. [Google Scholar] [CrossRef] [Green Version]
  61. Lai, L.; Leone, T.C.; Zechner, C.; Schaeffer, P.J.; Kelly, S.M.; Flanagan, D.P.; Medeiros, D.M.; Kovacs, A.; Kelly, D.P. Transcriptional coactivators PGC-1alpha and PGC-lbeta control overlapping programs required for perinatal maturation of the heart. Genes Dev. 2008, 22, 1948–1961. [Google Scholar] [CrossRef] [Green Version]
  62. Kadlec, A.O.; Chabowski, D.S.; Ait-Aissa, K.; Gutterman, D.D. Role of PGC-1α in Vascular Regulation: Implications for Atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 2016, 36, 1467–1474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Xue, Y.; Wei, Z.; Ding, H.; Wang, Q.; Zhou, Z.; Zheng, S.; Zhang, Y.; Hou, D.; Liu, Y.; Zen, K.; et al. MicroRNA-19b/221/222 induces endothelial cell dysfunction via suppression of PGC-1α in the progression of atherosclerosis. Atherosclerosis 2015, 241, 671–681. [Google Scholar] [CrossRef]
  64. Mccarthy, C.; Lieggi, N.T.; Barry, D.; Mooney, D.; de Gaetano, M.; James, W.G.; Mcclelland, S.; Barry, M.C.; Escoubet-Lozach, L.; Li, A.C.; et al. Macrophage PPAR gamma Co-activator-1 alpha participates in repressing foam cell formation and atherosclerosis in response to conjugated linoleic acid. EMBO Mol. Med. 2013, 5, 1443–1457. [Google Scholar] [CrossRef]
  65. Wei, Z.; Chong, H.; Jiang, Q.; Tang, Y.; Xu, J.; Wang, H.; Shi, Y.; Cui, L.; Li, J.; Zhang, Y.; et al. Smooth Muscle Overexpression of PGC1α Attenuates Atherosclerosis in Rabbits. Circ. Res. 2021, 129, E72–E86. [Google Scholar] [CrossRef] [PubMed]
  66. Stein, S.; Lohmann, C.; Handschin, C.; Stenfeldt, E.; Borén, J.; Lüscher, T.F.; Matter, C.M. ApoE−/− PGC-1α−/− mice display reduced IL-18 levels and do not develop enhanced atherosclerosis. PLoS ONE 2010, 5, 13539. [Google Scholar] [CrossRef] [Green Version]
  67. Xiong, S.; Patrushev, N.; Forouzandeh, F.; Hilenski, L.; Alexander, R.W. PGC-1α Modulates Telomere Function and DNA Damage in Protecting against Aging-Related Chronic Diseases. Cell Rep. 2015, 12, 1391–1399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Van Der Pluijm, I.; Burger, J.; Van Heijningen, P.M.; IJpma, A.; Van Vliet, N.; Milanese, C.; Schoonderwoerd, K.; Sluiter, W.; Ringuette, L.J.; Dekkers, D.H.W.; et al. Decreased mitochondrial respiration in aneurysmal aortas of Fibulin-4 mutant mice is linked to PGC1A regulation. Cardiovasc. Res. 2018, 114, 1776–1793. [Google Scholar] [CrossRef] [Green Version]
  69. Martin, O.J.; Lai, L.; Soundarapandian, M.M.; Leone, T.C.; Zorzano, A.; Keller, M.P.; Attie, A.D.; Muoio, D.M.; Kelly, D.P. A role for peroxisome proliferator-activated receptor γ coactivator-1 in the control of mitochondrial dynamics during postnatal cardiac growth. Circ. Res. 2014, 114, 626–636. [Google Scholar] [CrossRef] [Green Version]
  70. Arany, Z.; Novikov, M.; Chin, S.; Ma, Y.; Rosenzweig, A.; Spiegelman, B.M. Transverse aortic constriction leads to accelerated heart failure in mice lacking PPAR-gamma coactivator 1alpha. Proc. Natl. Acad. Sci. USA 2006, 103, 10086–10091. [Google Scholar] [CrossRef] [Green Version]
  71. Sawada, N.; Jiang, A.; Takizawa, F.; Safdar, A.; Manika, A.; Tesmenitsky, Y.; Kang, K.T.; Bischoff, J.; Kalwa, H.; Sartoretto, J.L.; et al. Endothelial PGC-1α mediates vascular dysfunction in diabetes. Cell Metab. 2014, 19, 246–258. [Google Scholar] [CrossRef] [Green Version]
  72. Dorn, G.W.; Vega, R.B.; Kelly, D.P. Mitochondrial biogenesis and dynamics in the developing and diseased heart. Genes Dev. 2015, 29, 1981–1991. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Finck, B.N.; Lehman, J.J.; Leone, T.C.; Welch, M.J.; Bennett, M.J.; Kovacs, A.; Han, X.; Gross, R.W.; Kozak, R.; Lopaschuk, G.D.; et al. The cardiac phenotype induced by PPARalpha overexpression mimics that caused by diabetes mellitus. J. Clin. Investig. 2002, 109, 121–130. [Google Scholar] [CrossRef] [PubMed]
  74. Watanabe, K.; Fujii, H.; Takahashi, T.; Kodama, M.; Aizawa, Y.; Ohta, Y.; Ono, T.; Hasegawa, G.; Naito, M.; Nakajima, T.; et al. Constitutive regulation of cardiac fatty acid metabolism through peroxisome proliferator-activated receptor alpha associated with age-dependent cardiac toxicity. J. Biol. Chem. 2000, 275, 22293–22299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Giguère, V. Transcriptional control of energy homeostasis by the estrogen-related receptors. Endocr. Rev. 2008, 29, 677–696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Huss, J.M.; Imahashi, K.; Dufour, C.R.; Weinheimer, C.J.; Courtois, M.; Kovacs, A.; Giguère, V.; Murphy, E.; Kelly, D.P. The nuclear receptor ERRalpha is required for the bioenergetic and functional adaptation to cardiac pressure overload. Cell Metab. 2007, 6, 25–37. [Google Scholar] [CrossRef] [Green Version]
  77. Alaynick, W.A.; Kondo, R.P.; Xie, W.; He, W.; Dufour, C.R.; Downes, M.; Jonker, J.W.W.; Giles, W.; Naviaux, R.K.; Giguère, V.; et al. ERRgamma directs and maintains the transition to oxidative metabolism in the postnatal heart. Cell Metab. 2007, 6, 13–24. [Google Scholar] [CrossRef] [Green Version]
  78. Scarpulla, R.C.; Vega, R.B.; Kelly, D.P. Transcriptional integration of mitochondrial biogenesis. Trends Endocrinol. Metab. 2012, 23, 459–466. [Google Scholar] [CrossRef] [Green Version]
  79. Gleyzer, N.; Vercauteren, K.; Scarpulla, R.C. Control of mitochondrial transcription specificity factors (TFB1M and TFB2M) by nuclear respiratory factors (NRF-1 and NRF-2) and PGC-1 family coactivators. Mol. Cell. Biol. 2005, 25, 1354–1366. [Google Scholar] [CrossRef] [Green Version]
  80. Huo, L.; Scarpulla, R.C. Mitochondrial DNA instability and peri-implantation lethality associated with targeted disruption of nuclear respiratory factor 1 in mice. Mol. Cell. Biol. 2001, 21, 644–654. [Google Scholar] [CrossRef] [Green Version]
  81. Ristevski, S.; O’Leary, D.A.; Thornell, A.P.; Owen, M.J.; Kola, I.; Hertzog, P.J. The ETS transcription factor GABPalpha is essential for early embryogenesis. Mol. Cell. Biol. 2004, 24, 5844–5849. [Google Scholar] [CrossRef] [Green Version]
  82. Li, F.; Wang, Y.; Zeller, K.I.; Potter, J.J.; Wonsey, D.R.; O’Donnell, K.A.; Kim, J.; Yustein, J.T.; Lee, L.A.; Dang, C.V. Myc stimulates nuclearly encoded mitochondrial genes and mitochondrial biogenesis. Mol. Cell. Biol. 2005, 25, 6225–6234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Ahuja, P.; Zhao, P.; Angelis, E.; Ruan, H.; Korge, P.; Olson, A.; Wang, Y.; Jin, E.S.; Jeffrey, F.M.; Portman, M.; et al. Myc controls transcriptional regulation of cardiac metabolism and mitochondrial biogenesis in response to pathological stress in mice. J. Clin. Investig. 2010, 120, 1494–1505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Campbell, C.T.; Kolesar, J.E.; Kaufman, B.A. Mitochondrial transcription factor A regulates mitochondrial transcription initiation, DNA packaging, and genome copy number. Biochim. Biophys. Acta 2012, 1819, 921–929. [Google Scholar] [CrossRef]
  85. Oller, J.; Gabandé-Rodríguez, E.; Ruiz-Rodríguez, M.J.; Desdín-Micó, G.; Aranda, J.F.; Rodrigues-Diez, R.; Ballesteros-Martínez, C.; Blanco, E.M.; Roldan-Montero, R.; Acuña, P.; et al. Extracellular Tuning of Mitochondrial Respiration Leads to Aortic Aneurysm. Circulation 2021, 143, 2091–2109. [Google Scholar] [CrossRef]
  86. Oller, J.; Gabandé-Rodríguez, E.; Roldan-Montero, R.; Ruiz-Rodríguez, M.J.; Redondo, J.M.; Martín-Ventura, J.L.; Mittelbrunn, M. Rewiring Vascular Metabolism Prevents Sudden Death due to Aortic Ruptures-Brief Report. Arterioscler. Thromb. Vasc. Biol. 2022, 42, 462–469. [Google Scholar] [CrossRef]
  87. Liesa, M.; Palacín, M.; Zorzano, A. Mitochondrial dynamics in mammalian health and disease. Physiol. Rev. 2009, 89, 799–845. [Google Scholar] [CrossRef] [Green Version]
  88. Kitada, M.; Koya, D. Renal protective effects of resveratrol. Oxid. Med. Cell. Longev. 2013, 2013, 568093. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Legros, F.; Lombès, A.; Frachon, P.; Rojo, M. Mitochondrial fusion in human cells is efficient, requires the inner membrane potential, and is mediated by mitofusins. Mol. Biol. Cell 2002, 13, 4343–4354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Olichon, A.; Emorine, L.J.; Descoins, E.; Pelloquin, L.; Brichese, L.; Gas, N.; Guillou, E.; Delettre, C.; Valette, A.; Hamel, C.P.; et al. The human dynamin-related protein OPA1 is anchored to the mitochondrial inner membrane facing the inter-membrane space. FEBS Lett. 2002, 523, 171–176. [Google Scholar] [CrossRef] [Green Version]
  91. Tilokani, L.; Nagashima, S.; Paupe, V.; Prudent, J. Mitochondrial dynamics: Overview of molecular mechanisms. Essays Biochem. 2018, 62, 341–360. [Google Scholar] [CrossRef] [Green Version]
  92. Losón, O.C.; Song, Z.; Chen, H.; Chan, D.C. Fis1, Mff, MiD49, and MiD51 mediate Drp1 recruitment in mitochondrial fission. Mol. Biol. Cell 2013, 24, 659–667. [Google Scholar] [CrossRef] [PubMed]
  93. Kuzmicic, J.; Del Campo, A.; López-Crisosto, C.; Morales, P.E.; Pennanen, C.; Bravo-Sagua, R.; Hechenleitner, J.; Zepeda, R.; Castro, P.F.; Verdejo, H.E.; et al. Mitochondrial dynamics: A potential new therapeutic target for heart failure. Rev. Esp. Cardiol. 2011, 64, 916–923. [Google Scholar] [CrossRef] [PubMed]
  94. Bach, D.; Naon, D.; Pich, S.; Soriano, F.X.; Vega, N.; Rieusset, J.; Laville, M.; Guillet, C.; Boirie, Y.; Wallberg-Henriksson, H.; et al. Expression of Mfn2, the Charcot-Marie-Tooth neuropathy type 2A gene, in human skeletal muscle: Effects of type 2 diabetes, obesity, weight loss, and the regulatory role of tumor necrosis factor alpha and interleukin-6. Diabetes 2005, 54, 2685–2693. [Google Scholar] [CrossRef] [Green Version]
  95. Soriano, F.X.; Liesa, M.; Bach, D.; Chan, D.C.; Palacín, M.; Zorzano, A. Evidence for a mitochondrial regulatory pathway defined by peroxisome proliferator-activated receptor-gamma coactivator-1 alpha, estrogen-related receptor-alpha, and mitofusin 2. Diabetes 2006, 55, 1783–1791. [Google Scholar] [CrossRef] [Green Version]
  96. MacVicar, T.; Langer, T. OPA1 processing in cell death and disease—The long and short of it. J. Cell Sci. 2016, 129, 2297–2306. [Google Scholar] [CrossRef] [Green Version]
  97. Tadato, B.; Heymann, J.A.W.; Song, Z.; Hinshaw, J.E.; Chan, D.C. OPA1 disease alleles causing dominant optic atrophy have defects in cardiolipin-stimulated GTP hydrolysis and membrane tubulation. Hum. Mol. Genet. 2010, 19, 2113–2122. [Google Scholar] [CrossRef] [Green Version]
  98. Weinberg, J.M. Mitochondrial biogenesis in kidney disease. J. Am. Soc. Nephrol. 2011, 22, 431–436. [Google Scholar] [CrossRef] [Green Version]
  99. Sharfuddin, A.A.; Molitoris, B.A. Pathophysiology of ischemic acute kidney injury. Nat. Rev. Nephrol. 2011, 7, 189–200. [Google Scholar] [CrossRef] [PubMed]
  100. Hoste, E.A.J.; De Corte, W. Clinical consequences of acute kidney injury. Contrib. Nephrol. 2011, 174, 56–64. [Google Scholar] [CrossRef]
  101. Vijayan, A. Tackling AKI: Prevention, timing of dialysis and follow-up. Nat. Rev. Nephrol. 2021, 17, 87–88. [Google Scholar] [CrossRef]
  102. Mitochondrial Bioenergetics during the Initiation of Mercuric Chloride-Induced Renal Injury. I. Direct Effects of In Vitro Mercuric Chloride on Renal Mitochondrial Function—PubMed. Available online: https://pubmed.ncbi.nlm.nih.gov/6458618/ (accessed on 19 March 2023).
  103. Ortega-Domínguez, B.; Aparicio-Trejo, O.E.; García-Arroyo, F.E.; León-Contreras, J.C.; Tapia, E.; Molina-Jijón, E.; Hernández-Pando, R.; Sánchez-Lozada, L.G.; Barrera-Oviedo, D.; Pedraza-Chaverri, J. Curcumin prevents cisplatin-induced renal alterations in mitochondrial bioenergetics and dynamic. Food Chem. Toxicol. 2017, 107, 373–385. [Google Scholar] [CrossRef] [PubMed]
  104. Gall, J.M.; Wang, Z.; Liesa, M.; Molina, A.; Havasi, A.; Schwartz, J.H.; Shirihai, O.; Borkan, S.C.; Bonegio, R.G.B. Role of mitofusin 2 in the renal stress response. PLoS ONE 2012, 7, 31074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Wang, Y.; Liu, Q.; Cai, J.; Wu, P.; Wang, D.; Shi, Y.; Huyan, T.; Su, J.; Li, X.; Wang, Q.; et al. Emodin prevents renal ischemia-reperfusion injury via suppression of CAMKII/DRP1-mediated mitochondrial fission. Eur. J. Pharmacol. 2022, 916, 174603. [Google Scholar] [CrossRef]
  106. Zsengellér, Z.K.; Ellezian, L.; Brown, D.; Horváth, B.; Mukhopadhyay, P.; Kalyanaraman, B.; Parikh, S.M.; Karumanchi, S.A.; Stillman, I.E.; Pacher, P. Cisplatin nephrotoxicity involves mitochondrial injury with impaired tubular mitochondrial enzyme activity. J. Histochem. Cytochem. 2012, 60, 521–529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Huang, Z.; Li, Q.; Yuan, Y.; Zhang, C.; Wu, L.; Liu, X.; Cao, W.; Guo, H.; Duan, S.; Xu, X.; et al. Renalase attenuates mitochondrial fission in cisplatin-induced acute kidney injury via modulating sirtuin-3. Life Sci. 2019, 222, 78–87. [Google Scholar] [CrossRef] [PubMed]
  108. Sun, L.; Yuan, Q.; Xu, T.; Yao, L.; Feng, J.; Ma, J.; Wang, L.; Lu, C.; Wang, D. Pioglitazone Improves Mitochondrial Function in the Remnant Kidney and Protects against Renal Fibrosis in 5/6 Nephrectomized Rats. Front. Pharmacol. 2017, 8, 545. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Wang, Y.; Lu, M.; Xiong, L.; Fan, J.; Zhou, Y.; Li, H.; Peng, X.; Zhong, Z.; Wang, Y.; Huang, F.; et al. Drp1-mediated mitochondrial fission promotes renal fibroblast activation and fibrogenesis. Cell Death Dis. 2020, 11, 29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Li, S.; Lin, Q.; Shao, X.; Zhu, X.; Wu, J.; Wu, B.; Zhang, M.; Zhou, W.; Zhou, Y.; Jin, H.; et al. Drp1-regulated PARK2-dependent mitophagy protects against renal fibrosis in unilateral ureteral obstruction. Free Radic. Biol. Med. 2020, 152, 632–649. [Google Scholar] [CrossRef]
  111. Zhan, M.; Usman, I.; Yu, J.; Ruan, L.; Bian, X.; Yang, J.; Yang, S.; Sun, L.; Kanwar, Y.S. Perturbations in mitochondrial dynamics by p66Shc lead to renal tubular oxidative injury in human diabetic nephropathy. Clin. Sci. 2018, 132, 1297–1314. [Google Scholar] [CrossRef]
  112. Tagaya, M.; Kume, S.; Yasuda-Yamahara, M.; Kuwagata, S.; Yamahara, K.; Takeda, N.; Tanaka, Y.; Chin-Kanasaki, M.; Nakae, Y.; Yokoi, H.; et al. Inhibition of mitochondrial fission protects podocytes from albumin-induced cell damage in diabetic kidney disease. Biochim. Biophys. Acta Mol. Basis Dis. 2022, 1868, 166368. [Google Scholar] [CrossRef]
  113. Ayanga, B.A.; Badal, S.S.; Wang, Y.; Galvan, D.L.; Chang, B.H.; Schumacker, P.T.; Danesh, F.R. Dynamin-Related Protein 1 Deficiency Improves Mitochondrial Fitness and Protects against Progression of Diabetic Nephropathy. J. Am. Soc. Nephrol. 2016, 27, 2733–2747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Qin, X.; Zhao, Y.; Gong, J.; Huang, W.; Su, H.; Yuan, F.; Fang, K.; Wang, D.; Li, J.; Zou, X.; et al. Berberine Protects Glomerular Podocytes via Inhibiting Drp1-Mediated Mitochondrial Fission and Dysfunction. Theranostics 2019, 9, 1698–1713. [Google Scholar] [CrossRef]
  115. Forte, M.; Schirone, L.; Ameri, P.; Basso, C.; Catalucci, D.; Modica, J.; Chimenti, C.; Crotti, L.; Frati, G.; Rubattu, S.; et al. The role of mitochondrial dynamics in cardiovascular diseases. Br. J. Pharmacol. 2021, 178, 2060–2076. [Google Scholar] [CrossRef] [PubMed]
  116. Xia, Y.; Zhang, X.; An, P.; Luo, J.; Luo, Y. Mitochondrial Homeostasis in VSMCs as a Central Hub in Vascular Remodeling. Int. J. Mol. Sci. 2023, 24, 3483. [Google Scholar] [CrossRef]
  117. Alavi, M.V.; Bette, S.; Schimpf, S.; Schuettauf, F.; Schraermeyer, U.; Wehrl, H.F.; Ruttiger, L.; Beck, S.C.; Tonagel, F.; Pichler, B.J.; et al. A splice site mutation in the murine Opa1 gene features pathology of autosomal dominant optic atrophy. Brain 2007, 130, 1029–1042. [Google Scholar] [CrossRef] [Green Version]
  118. Chen, H.; Detmer, S.A.; Ewald, A.J.; Griffin, E.E.; Fraser, S.E.; Chan, D.C. Mitofusins Mfn1 and Mfn2 coordinately regulate mitochondrial fusion and are essential for embryonic development. J. Cell Biol. 2003, 160, 189–200. [Google Scholar] [CrossRef] [PubMed]
  119. Chen, Y.; Liu, Y.; Dorn, G.W. Mitochondrial fusion is essential for organelle function and cardiac homeostasis. Circ. Res. 2011, 109, 1327–1331. [Google Scholar] [CrossRef]
  120. Song, M.; Mihara, K.; Chen, Y.; Scorrano, L.; Dorn, G.W. Mitochondrial fission and fusion factors reciprocally orchestrate mitophagic culling in mouse hearts and cultured fibroblasts. Cell Metab. 2015, 21, 273–286. [Google Scholar] [CrossRef] [Green Version]
  121. Kasahara, A.; Cipolat, S.; Chen, Y.; Dorn, G.W.; Scorrano, L. Mitochondrial fusion directs cardiomyocyte differentiation via calcineurin and Notch signaling. Science 2013, 342, 734–737. [Google Scholar] [CrossRef] [Green Version]
  122. Papanicolaou, K.N.; Ngoh, G.A.; Dabkowski, E.R.; O’Connell, K.A.; Ribeiro, R.F.; Stanley, W.C.; Walsh, K. Cardiomyocyte deletion of mitofusin-1 leads to mitochondrial fragmentation and improves tolerance to ROS-induced mitochondrial dysfunction and cell death. Am. J. Physiol. Heart Circ. Physiol. 2012, 302, H167–H179. [Google Scholar] [CrossRef] [Green Version]
  123. Chen, Y.; Dorn, G.W. PINK1-phosphorylated mitofusin 2 is a Parkin receptor for culling damaged mitochondria. Science 2013, 340, 471–475. [Google Scholar] [CrossRef] [Green Version]
  124. Papanicolaou, K.N.; Khairallah, R.J.; Ngoh, G.A.; Chikando, A.; Luptak, I.; O’Shea, K.M.; Riley, D.D.; Lugus, J.J.; Colucci, W.S.; Lederer, W.J.; et al. Mitofusin-2 maintains mitochondrial structure and contributes to stress-induced permeability transition in cardiac myocytes. Mol. Cell. Biol. 2011, 31, 1309–1328. [Google Scholar] [CrossRef] [Green Version]
  125. De Brito, O.M.; Scorrano, L. Mitofusin 2 tethers endoplasmic reticulum to mitochondria. Nature 2008, 456, 605–610. [Google Scholar] [CrossRef]
  126. Le Page, S.; Niro, M.; Fauconnier, J.; Cellier, L.; Tamareille, S.; Gharib, A.; Chevrollier, A.; Loufrani, L.; Grenier, C.; Kamel, R.; et al. Increase in Cardiac Ischemia-Reperfusion Injuries in Opa1+/− Mouse Model. PLoS ONE 2016, 11, e0164066. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Ishihara, T.; Ban-Ishihara, R.; Maeda, M.; Matsunaga, Y.; Ichimura, A.; Kyogoku, S.; Aoki, H.; Katada, S.; Nakada, K.; Nomura, M.; et al. Dynamics of mitochondrial DNA nucleoids regulated by mitochondrial fission is essential for maintenance of homogeneously active mitochondria during neonatal heart development. Mol. Cell. Biol. 2015, 35, 211–223. [Google Scholar] [CrossRef] [Green Version]
  128. Kageyama, Y.; Hoshijima, M.; Seo, K.; Bedja, D.; Sysa-Shah, P.; Andrabi, S.A.; Chen, W.; Höke, A.; Dawson, V.L.; Dawson, T.M.; et al. Parkin-independent mitophagy requires Drp1 and maintains the integrity of mammalian heart and brain. EMBO J. 2014, 33, 2798–2813. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Song, M.; Gong, G.; Burelle, Y.; Gustafsson, Å.B.; Kitsis, R.N.; Matkovich, S.J.; Dorn, G.W. Interdependence of Parkin-Mediated Mitophagy and Mitochondrial Fission in Adult Mouse Hearts. Circ. Res. 2015, 117, 346–351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Ikeda, Y.; Shirakabe, A.; Maejima, Y.; Zhai, P.; Sciarretta, S.; Toli, J.; Nomura, M.; Mihara, K.; Egashira, K.; Ohishi, M.; et al. Endogenous Drp1 mediates mitochondrial autophagy and protects the heart against energy stress. Circ. Res. 2015, 116, 264–278. [Google Scholar] [CrossRef] [PubMed]
  131. Chaanine, A.H.; Kohlbrenner, E.; Gamb, S.I.; Guenzel, A.J.; Klaus, K.; Fayyaz, A.U.; Sreekumaran Nair, K.; Hajjar, R.J.; Redfield, M.M. FOXO3a regulates BNIP3 and modulates mitochondrial calcium, dynamics, and function in cardiac stress. Am. J. Physiol. Heart Circ. Physiol. 2016, 311, H1540–H1559. [Google Scholar] [CrossRef] [Green Version]
  132. Montaigne, D.; Marechal, X.; Coisne, A.; Debry, N.; Modine, T.; Fayad, G.; Potelle, C.; El Arid, J.M.; Mouton, S.; Sebti, Y.; et al. Myocardial contractile dysfunction is associated with impaired mitochondrial function and dynamics in type 2 diabetic but not in obese patients. Circulation 2014, 130, 554–564. [Google Scholar] [CrossRef] [Green Version]
  133. Chen, K.H.; Guo, X.; Ma, D.; Guo, Y.; Li, Q.; Yang, D.; Li, P.; Qiu, X.; Wen, S.; Xiao, R.P.; et al. Dysregulation of HSG triggers vascular proliferative disorders. Nat. Cell Biol. 2004, 6, 872–883. [Google Scholar] [CrossRef] [PubMed]
  134. Liu, C.; Ge, B.; He, C.; Zhang, Y.; Liu, X.; Liu, K.; Qian, C.; Zhang, Y.; Peng, W.; Guo, X. Mitofusin 2 decreases intracellular lipids in macrophages by regulating peroxisome proliferator-activated receptor-γ. Biochem. Biophys. Res. Commun. 2014, 450, 500–506. [Google Scholar] [CrossRef] [PubMed]
  135. Guo, Y.; Chen, K.; Gao, W.; Li, Q.; Chen, L.; Wang, G.; Tang, J. Overexpression of Mitofusin 2 inhibited oxidized low-density lipoprotein induced vascular smooth muscle cell proliferation and reduced atherosclerotic lesion formation in rabbit. Biochem. Biophys. Res. Commun. 2007, 363, 411–417. [Google Scholar] [CrossRef] [PubMed]
  136. Bi, X.; Du, C.; Wang, X.; Wang, X.Y.; Han, W.; Wang, Y.; Qiao, Y.; Zhu, Y.; Ran, L.; Liu, Y.; et al. Mitochondrial Damage-Induced Innate Immune Activation in Vascular Smooth Muscle Cells Promotes Chronic Kidney Disease-Associated Plaque Vulnerability. Adv. Sci. 2021, 8, 2002738. [Google Scholar] [CrossRef]
  137. Salabei, J.K.; Hill, B.G. Mitochondrial fission induced by platelet-derived growth factor regulates vascular smooth muscle cell bioenergetics and cell proliferation. Redox Biol. 2013, 1, 542–551. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Lim, S.; Lee, S.Y.; Seo, H.H.; Ham, O.; Lee, C.; Park, J.H.; Lee, J.; Seung, M.; Yun, I.; Han, S.M.; et al. Regulation of mitochondrial morphology by positive feedback interaction between PKCδ and Drp1 in vascular smooth muscle cell. J. Cell. Biochem. 2015, 116, 648–660. [Google Scholar] [CrossRef]
  139. 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]
  140. Wang, Q.; Zhang, M.; Torres, G.; Wu, S.; Ouyang, C.; Xie, Z.; Zou, M.H. Metformin Suppresses Diabetes-Accelerated Atherosclerosis via the Inhibition of Drp1-Mediated Mitochondrial Fission. Diabetes 2017, 66, 193–205. [Google Scholar] [CrossRef] [Green Version]
  141. Rogers, M.A.; Maldonado, N.; Hutcheson, J.D.; Goettsch, C.; Goto, S.; Yamada, I.; Faits, T.; Sesaki, H.; Aikawa, M.; Aikawa, E. Dynamin-Related Protein 1 Inhibition Attenuates Cardiovascular Calcification in the Presence of Oxidative Stress. Circ. Res. 2017, 121, 220–233. [Google Scholar] [CrossRef]
  142. Peng, C.; Rao, W.; Zhang, L.; Wang, K.; Hui, H.; Wang, L.; Su, N.; Luo, P.; Hao, Y.L.; Tu, Y.; et al. Mitofusin 2 ameliorates hypoxia-induced apoptosis via mitochondrial function and signaling pathways. Int. J. Biochem. Cell Biol. 2015, 69, 29–40. [Google Scholar] [CrossRef]
  143. Barsoum, M.J.; Yuan, H.; Gerencser, A.A.; Liot, G.; Kushnareva, Y.; Gräber, S.; Kovacs, I.; Lee, W.D.; Waggoner, J.; Cui, J.; et al. Nitric oxide-induced mitochondrial fission is regulated by dynamin-related GTPases in neurons. EMBO J. 2006, 25, 3900–3911. [Google Scholar] [CrossRef] [PubMed]
  144. Grohm, J.; Kim, S.-W.; Mamrak, U.; Tobaben, S.; Cassidy-Stone, A.; Nunnari, J.; Plesnila, N.; Culmsee, C. Inhibition of Drp1 provides neuroprotection in vitro and in vivo. Cell Death Differ. 2012, 19, 1446–1458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Zhao, T.; Huang, X.; Han, L.; Wang, X.; Cheng, H.; Zhao, Y.; Chen, Q.; Chen, J.; Cheng, H.; Xiao, R.; et al. Central role of mitofusin 2 in autophagosome-lysosome fusion in cardiomyocytes. J. Biol. Chem. 2012, 287, 23615–23625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Cooper, H.A.; Cicalese, S.; Preston, K.J.; Kawai, T.; Okuno, K.; Choi, E.T.; Kasahara, S.; Uchida, H.A.; Otaka, N.; Scalia, R.; et al. Targeting mitochondrial fission as a potential therapeutic for abdominal aortic aneurysm. Cardiovasc. Res. 2021, 117, 971–982. [Google Scholar] [CrossRef]
  147. Wang, X.; Li, S.; Liu, L.; Jian, Z.; Cui, T.; Yang, Y.; Guo, S.; Yi, X.; Wang, G.; Li, C.; et al. Role of the aryl hydrocarbon receptor signaling pathway in promoting mitochondrial biogenesis against oxidative damage in human melanocytes. J. Dermatol. Sci. 2019, 96, 33–41. [Google Scholar] [CrossRef] [PubMed]
  148. Ray, P.D.; Huang, B.W.; Tsuji, Y. Reactive oxygen species (ROS) homeostasis and redox regulation in cellular signaling. Cell. Signal. 2012, 24, 981–990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Wu, S.; Zhou, F.; Zhang, Z.; Xing, D. Mitochondrial oxidative stress causes mitochondrial fragmentation via differential modulation of mitochondrial fission-fusion proteins. FEBS J. 2011, 278, 941–954. [Google Scholar] [CrossRef]
  150. Aparicio-Trejo, O.E.; Reyes-Fermín, L.M.; Briones-Herrera, A.; Tapia, E.; León-Contreras, J.C.; Hernández-Pando, R.; Sánchez-Lozada, L.G.; Pedraza-Chaverri, J. Protective effects of N-acetyl-cysteine in mitochondria bioenergetics, oxidative stress, dynamics and S-glutathionylation alterations in acute kidney damage induced by folic acid. Free Radic. Biol. Med. 2019, 130, 379–396. [Google Scholar] [CrossRef]
  151. Aparicio-Trejo, O.E.; Avila-Rojas, S.H.; Tapia, E.; Rojas-Morales, P.; León-Contreras, J.C.; Martínez-Klimova, E.; Hernández-Pando, R.; Sánchez-Lozada, L.G.; Pedraza-Chaverri, J. Chronic impairment of mitochondrial bioenergetics and β-oxidation promotes experimental AKI-to-CKD transition induced by folic acid. Free Radic. Biol. Med. 2020, 154, 18–32. [Google Scholar] [CrossRef]
  152. Mukhopadhyay, P.; Horváth, B.; Zsengellér, Z.; Zielonka, J.; Tanchian, G.; Holovac, E.; Kechrid, M.; Patel, V.; Stillman, I.E.; Parikh, S.M.; et al. Mitochondrial-targeted antioxidants represent a promising approach for prevention of cisplatin-induced nephropathy. Free Radic. Biol. Med. 2012, 52, 497–506. [Google Scholar] [CrossRef] [Green Version]
  153. Gong, X.; Duan, Y.; Zheng, J.; Ye, Z.; Hei, T.K. Tetramethylpyrazine Prevents Contrast-Induced Nephropathy via Modulating Tubular Cell Mitophagy and Suppressing Mitochondrial Fragmentation, CCL2/CCR2-Mediated Inflammation, and Intestinal Injury. Oxid. Med. Cell. Longev. 2019, 2019, 7096912. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Regan, J.D.; Setlow, R.B.; Ley, R.D. Normal and defective repair of damaged DNA in human cells: A sensitive assay utilizing the photolysis of bromodeoxyuridine. Proc. Natl. Acad. Sci. USA 1971, 68, 708–712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Plotnikov, E.Y.; Chupyrkina, A.A.; Jankauskas, S.S.; Pevzner, I.B.; Silachev, D.N.; Skulachev, V.P.; Zorov, D.B. Mechanisms of nephroprotective effect of mitochondria-targeted antioxidants under rhabdomyolysis and ischemia/reperfusion. Biochim. Biophys. Acta 2011, 1812, 77–86. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Pabla, N.; Bajwa, A. Role of Mitochondrial Therapy for Ischemic-Reperfusion Injury and Acute Kidney Injury. Nephron 2022, 146, 253–258. [Google Scholar] [CrossRef]
  157. Jankauskas, S.S.; Silachev, D.N.; Andrianova, N.V.; Pevzner, I.B.; Zorova, L.D.; Popkov, V.A.; Plotnikov, E.Y.; Zorov, D.B. Aged kidney: Can we protect it? Autophagy, mitochondria and mechanisms of ischemic preconditioning. Cell Cycle 2018, 17, 1291–1309. [Google Scholar] [CrossRef] [Green Version]
  158. Zhao, M.; Wang, Y.; Li, L.; Liu, S.; Wang, C.; Yuan, Y.; Yang, G.; Chen, Y.; Cheng, J.; Lu, Y.; et al. Mitochondrial ROS promote mitochondrial dysfunction and inflammation in ischemic acute kidney injury by disrupting TFAM-mediated mtDNA maintenance. Theranostics 2021, 11, 1845–1863. [Google Scholar] [CrossRef]
  159. Maekawa, H.; Inoue, T.; Ouchi, H.; Jao, T.M.; Inoue, R.; Nishi, H.; Fujii, R.; Ishidate, F.; Tanaka, T.; Tanaka, Y.; et al. Mitochondrial Damage Causes Inflammation via cGAS-STING Signaling in Acute Kidney Injury. Cell Rep. 2019, 29, 1261–1273.e6. [Google Scholar] [CrossRef] [Green Version]
  160. Lood, C.; Blanco, L.P.; Purmalek, M.M.; Carmona-Rivera, C.; De Ravin, S.S.; Smith, C.K.; Malech, H.L.; Ledbetter, J.A.; Elkon, K.B.; Kaplan, M.J. Neutrophil extracellular traps enriched in oxidized mitochondrial DNA are interferogenic and contribute to lupus-like disease. Nat. Med. 2016, 22, 146–153. [Google Scholar] [CrossRef] [Green Version]
  161. Mohammad, R.S.; Lokhandwala, M.F.; Banday, A.A. Age-Related Mitochondrial Impairment and Renal Injury Is Ameliorated by Sulforaphane via Activation of Transcription Factor NRF2. Antioxidants 2022, 11, 156. [Google Scholar] [CrossRef]
  162. Marquez-Exposito, L.; Tejedor-Santamaria, L.; Valentijn, F.A.; Tejera-Muñoz, A.; Rayego-Mateos, S.; Marchant, V.; Rodrigues-Diez, R.R.; Rubio-Soto, I.; Knoppert, S.N.; Ortiz, A.; et al. Oxidative Stress and Cellular Senescence Are Involved in the Aging Kidney. Antioxidants 2022, 11, 301. [Google Scholar] [CrossRef]
  163. Hu, X.; Yang, Y.; Tang, S.; Chen, Q.; Zhang, M.; Ma, J.; Qin, J.; Yu, H. Anti-Aging Effects of Anthocyanin Extracts of Sambucus canadensis Caused by Targeting Mitochondrial-Induced Oxidative Stress. Int. J. Mol. Sci. 2023, 24, 1528. [Google Scholar] [CrossRef]
  164. Kolosova, N.G.; Kozhevnikova, O.S.; Muraleva, N.A.; Rudnitskaya, E.A.; Rumyantseva, Y.V.; Stefanova, N.A.; Telegina, D.V.; Tyumentsev, M.A.; Fursova, A.Z. SkQ1 as a Tool for Controlling Accelerated Senescence Program: Experiments with OXYS Rats. Biochemistry 2022, 87, 1552–1562. [Google Scholar] [CrossRef]
  165. Szeto, H.H.; Liu, S.; Soong, Y.; Alam, N.; Prusky, G.T.; Seshan, S.V. Protection of mitochondria prevents high-fat diet-induced glomerulopathy and proximal tubular injury. Kidney Int. 2016, 90, 997–1011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Szeto, H.H.; Liu, S.; Soong, Y.; Wu, D.; Darrah, S.F.; Cheng, F.Y.; Zhao, Z.; Ganger, M.; Tow, C.Y.; Seshan, S.V. Mitochondria-targeted peptide accelerates ATP recovery and reduces ischemic kidney injury. J. Am. Soc. Nephrol. 2011, 22, 1041–1052. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Hou, Y.; Li, S.; Wu, M.; Wei, J.; Ren, Y.; Du, C.; Wu, H.; Han, C.; Duan, H.; Shi, Y. Mitochondria-targeted peptide SS-31 attenuates renal injury via an antioxidant effect in diabetic nephropathy. Am. J. Physiol. Ren. Physiol. 2016, 310, F547–F559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Saad, A.; Herrmann, S.M.S.; Eirin, A.; Ferguson, C.M.; Glockner, J.F.; Bjarnason, H.; McKusick, M.A.; Misra, S.; Lerman, L.O.; Textor, S.C. Phase 2a Clinical Trial of Mitochondrial Protection (Elamipretide) During Stent Revascularization in Patients With Atherosclerotic Renal Artery Stenosis. Circ. Cardiovasc. Interv. 2017, 10, 5487. [Google Scholar] [CrossRef] [PubMed]
  169. Buchke, S.; Sharma, M.; Bora, A.; Relekar, M.; Bhanu, P.; Kumar, J. Mitochondria-Targeted, Nanoparticle-Based Drug-Delivery Systems: Therapeutics for Mitochondrial Disorders. Life 2022, 12, 657. [Google Scholar] [CrossRef] [PubMed]
  170. Kwon, H.J.; Cha, M.Y.; Kim, D.; Kim, D.K.; Soh, M.; Shin, K.; Hyeon, T.; Mook-Jung, I. Mitochondria-Targeting Ceria Nanoparticles as Antioxidants for Alzheimer’s Disease. ACS Nano 2016, 10, 2860–2870. [Google Scholar] [CrossRef] [PubMed]
  171. Cheng, Y.; Liu, D.; Zhang, C.; Cui, H.; Liu, M.; Zhang, B.; Mei, Q.; Lu, Z.; Zhou, S. Mitochondria-targeted antioxidant delivery for precise treatment of myocardial ischemia-reperfusion injury through a multistage continuous targeted strategy. Nanomedicine 2019, 16, 236–249. [Google Scholar] [CrossRef] [PubMed]
  172. Lozano, O.; Lázaro-Alfaro, A.; Silva-Platas, C.; Oropeza-Almazán, Y.; Torres-Quintanilla, A.; Bernal-Ramírez, J.; Alves-Figueiredo, H.; García-Rivas, G. Nanoencapsulated Quercetin Improves Cardioprotection during Hypoxia-Reoxygenation Injury through Preservation of Mitochondrial Function. Oxid. Med. Cell. Longev. 2019, 2019, 7683051. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Sarkar, S.; Mukherjee, A.; Das, N.; Swarnakar, S. Protective roles of nanomelatonin in cerebral ischemia-reperfusion of aged brain: Matrixmetalloproteinases as regulators. Exp. Gerontol. 2017, 92, 13–22. [Google Scholar] [CrossRef]
  174. Quan, Y.; Xin, Y.; Tian, G.; Zhou, J.; Liu, X. Mitochondrial ROS-Modulated mtDNA: A Potential Target for Cardiac Aging. Oxid. Med. Cell. Longev. 2020, 2020, 9423593. [Google Scholar] [CrossRef] [Green Version]
  175. Marin, W.; Marin, D.; Ao, X.; Liu, Y. Mitochondria as a therapeutic target for cardiac ischemia-reperfusion injury (Review). Int. J. Mol. Med. 2021, 47, 485–499. [Google Scholar] [CrossRef]
  176. Ungvari, Z.; Tarantini, S.; Donato, A.J.; Galvan, V.; Csiszar, A. Mechanisms of Vascular Aging. Circ. Res. 2018, 123, 849–867. [Google Scholar] [CrossRef] [PubMed]
  177. Nakada, Y.; Canseco, D.C.; Thet, S.; Abdisalaam, S.; Asaithamby, A.; Santos, C.X.; Shah, A.M.; Zhang, H.; Faber, J.E.; Kinter, M.T.; et al. Hypoxia induces heart regeneration in adult mice. Nature 2017, 541, 222–227. [Google Scholar] [CrossRef] [PubMed]
  178. Tsushima, K.; Bugger, H.; Wende, A.R.; Soto, J.; Jenson, G.A.; Tor, A.R.; McGlauflin, R.; Kenny, H.C.; Zhang, Y.; Souvenir, R.; et al. Mitochondrial Reactive Oxygen Species in Lipotoxic Hearts Induce Post-Translational Modifications of AKAP121, DRP1, and OPA1 That Promote Mitochondrial Fission. Circ. Res. 2018, 122, 58–73. [Google Scholar] [CrossRef] [PubMed]
  179. Brand, M.D.; Goncalves, R.L.S.; Orr, A.L.; Vargas, L.; Gerencser, A.A.; Borch Jensen, M.; Wang, Y.T.; Melov, S.; Turk, C.N.; Matzen, J.T.; et al. Suppressors of Superoxide-H2O2 Production at Site IQ of Mitochondrial Complex I Protect against Stem Cell Hyperplasia and Ischemia-Reperfusion Injury. Cell Metab. 2016, 24, 582–592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Dai, D.F.; Santana, L.F.; Vermulst, M.; Tomazela, D.M.; Emond, M.J.; MacCoss, M.J.; Gollahon, K.; Martin, G.M.; Loeb, L.A.; Ladiges, W.C.; et al. Overexpression of catalase targeted to mitochondria attenuates murine cardiac aging. Circulation 2009, 119, 2789–2797. [Google Scholar] [CrossRef]
  181. Dai, D.F.; Rabinovitch, P. Mitochondrial oxidative stress mediates induction of autophagy and hypertrophy in angiotensin-II treated mouse hearts. Autophagy 2011, 7, 917–918. [Google Scholar] [CrossRef] [Green Version]
  182. Luptak, I.; Qin, F.; Sverdlov, A.L.; Pimentel, D.R.; Panagia, M.; Croteau, D.; Siwik, D.A.; Bachschmid, M.M.; He, H.; Balschi, J.A.; et al. Energetic Dysfunction Is Mediated by Mitochondrial Reactive Oxygen Species and Precedes Structural Remodeling in Metabolic Heart Disease. Antioxid. Redox Signal. 2019, 31, 539–549. [Google Scholar] [CrossRef]
  183. Lee, C.F.; Liu, C.Y.; Hsieh, R.H.; Wei, Y.H. Oxidative stress-induced depolymerization of microtubules and alteration of mitochondrial mass in human cells. Ann. N. Y. Acad. Sci. 2005, 1042, 246–254. [Google Scholar] [CrossRef]
  184. Zhang, Y.; Guallar, E.; Ashar, F.N.; Longchamps, R.J.; Castellani, C.A.; Lane, J.; Grove, M.L.; Coresh, J.; Sotoodehnia, N.; Ilkhanoff, L.; et al. Association between mitochondrial DNA copy number and sudden cardiac death: Findings from the Atherosclerosis Risk in Communities study (ARIC). Eur. Heart J. 2017, 38, 3443–3448. [Google Scholar] [CrossRef] [Green Version]
  185. Ashar, F.N.; Zhang, Y.; Longchamps, R.J.; Lane, J.; Moes, A.; Grove, M.L.; Mychaleckyj, J.C.; Taylor, K.D.; Coresh, J.; Rotter, J.I.; et al. Association of Mitochondrial DNA Copy Number With Cardiovascular Disease. JAMA Cardiol. 2017, 2, 1247–1255. [Google Scholar] [CrossRef] [PubMed]
  186. Peng, W.; Cai, G.; Xia, Y.; Chen, J.; Wu, P.; Wang, Z.; Li, G.; Wei, D. Mitochondrial Dysfunction in Atherosclerosis. DNA Cell Biol. 2019, 38, 597–606. [Google Scholar] [CrossRef]
  187. Gray, K.; Kumar, S.; Figg, N.; Harrison, J.; Baker, L.; Mercer, J.; Littlewood, T.; Bennett, M. Effects of DNA damage in smooth muscle cells in atherosclerosis. Circ. Res. 2015, 116, 816–826. [Google Scholar] [CrossRef] [PubMed]
  188. Ballinger, S.W.; Patterson, C.; Knight-Lozano, C.A.; Burow, D.L.; Conklin, C.A.; Hu, Z.; Reuf, J.; Horaist, C.; Lebovitz, R.; Hunter, G.C.; et al. Mitochondrial integrity and function in atherogenesis. Circulation 2002, 106, 544–549. [Google Scholar] [CrossRef]
  189. Wei, Y.; Corbalán-Campos, J.; Gurung, R.; Natarelli, L.; Zhu, M.; Exner, N.; Erhard, F.; Greulich, F.; Geißler, C.; Uhlenhaut, N.H.; et al. Dicer in Macrophages Prevents Atherosclerosis by Promoting Mitochondrial Oxidative Metabolism. Circulation 2018, 138, 2007–2020. [Google Scholar] [CrossRef] [Green Version]
  190. Kujoth, C.C.; Hiona, A.; Pugh, T.D.; Someya, S.; Panzer, K.; Wohlgemuth, S.E.; Hofer, T.; Seo, A.Y.; Sullivan, R.; Jobling, W.A.; et al. Mitochondrial DNA mutations, oxidative stress, and apoptosis in mammalian aging. Science 2005, 309, 481–484. [Google Scholar] [CrossRef] [PubMed]
  191. Trifunovic, A.; Hansson, A.; Wredenberg, A.; Rovio, A.T.; Dufour, E.; Khvorostov, I.; Spelbrink, J.N.; Wibom, R.; Jacobs, H.T.; Larsson, N.G. Somatic mtDNA mutations cause aging phenotypes without affecting reactive oxygen species production. Proc. Natl. Acad. Sci. USA 2005, 102, 17993–17998. [Google Scholar] [CrossRef] [Green Version]
  192. Mercer, J.R.; Yu, E.; Figg, N.; Cheng, K.K.; Prime, T.A.; Griffin, J.L.; Masoodi, M.; Vidal-Puig, A.; Murphy, M.P.; Bennett, M.R. The mitochondria-targeted antioxidant MitoQ decreases features of the metabolic syndrome in ATM+/−/ApoE−/− mice. Free Radic. Biol. Med. 2012, 52, 841–849. [Google Scholar] [CrossRef]
  193. Wang, X.; Cui, N.; Zhang, S.; Liu, Z.; Ma, J.; Ming, L. Leukocyte telomere length, mitochondrial DNA copy number, and coronary artery disease risk and severity: A two-stage case-control study of 3064 Chinese subjects. Atherosclerosis 2019, 284, 165–172. [Google Scholar] [CrossRef]
  194. Zhang, D.; Li, Y.; Heims-Waldron, D.; Bezzerides, V.; Guatimosim, S.; Guo, Y.; Gu, F.; Zhou, P.; Lin, Z.; Ma, Q.; et al. Mitochondrial Cardiomyopathy Caused by Elevated Reactive Oxygen Species and Impaired Cardiomyocyte Proliferation. Circ. Res. 2018, 122, 74–87. [Google Scholar] [CrossRef]
  195. Lesnefsky, E.J.; Chen, Q.; Tandler, B.; Hoppel, C.L. Mitochondrial Dysfunction and Myocardial Ischemia-Reperfusion: Implications for Novel Therapies. Annu. Rev. Pharmacol. Toxicol. 2017, 57, 535–565. [Google Scholar] [CrossRef]
  196. Reily, C.; Mitchell, T.; Chacko, B.K.; Benavides, G.A.; Murphy, M.P.; Darley-Usmar, V.M. Mitochondrially targeted compounds and their impact on cellular bioenergetics. Redox Biol. 2013, 1, 86–93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Adlam, V.J.; Harrison, J.C.; Porteous, C.M.; James, A.M.; Smith, R.A.J.; Murphy, M.P.; Sammut, I.A. Targeting an antioxidant to mitochondria decreases cardiac ischemia-reperfusion injury. FASEB J. 2005, 19, 1088–1095. [Google Scholar] [CrossRef]
  198. Zhang, Z.W.; Xu, X.C.; Liu, T.; Yuan, S. Mitochondrion-Permeable Antioxidants to Treat ROS-Burst-Mediated Acute Diseases. Oxid. Med. Cell. Longev. 2016, 2016, 6859523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Lee, F.Y.; Shao, P.L.; Wallace, C.G.; Chua, S.; Sung, P.H.; Ko, S.F.; Chai, H.T.; Chung, S.Y.; Chen, K.H.; Lu, H.I.; et al. Combined Therapy with SS31 and Mitochondria Mitigates Myocardial Ischemia-Reperfusion Injury in Rats. Int. J. Mol. Sci. 2018, 19, 2782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Baguisi, A.; Casale, R.A.; Kates, S.A.; Lader, A.S.; Stewart, K.; Beeuwkes, R. CMX-2043 Efficacy in a Rat Model of Cardiac Ischemia-Reperfusion Injury. J. Cardiovasc. Pharmacol. Ther. 2016, 21, 563–569. [Google Scholar] [CrossRef] [PubMed]
  201. Esplugues, J.V.; Rocha, M.; Nuñez, C.; Bosca, I.; Ibiza, S.; Herance, J.R.; Ortega, A.; Serrador, J.M.; D’Ocon, P.; Victor, V.M. Complex I dysfunction and tolerance to nitroglycerin: An approach based on mitochondrial-targeted antioxidants. Circ. Res. 2006, 99, 1067–1075. [Google Scholar] [CrossRef] [Green Version]
  202. Graham, D.; Huynh, N.N.; Hamilton, C.A.; Beattie, E.; Smith, R.A.J.; Cochemé, H.M.; Murphy, M.P.; Dominiczak, A.F. Mitochondria-targeted antioxidant MitoQ10 improves endothelial function and attenuates cardiac hypertrophy. Hypertens 2009, 54, 322–328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. van Empel, V.P.M.; Bertrand, A.T.; van Oort, R.J.; van der Nagel, R.; Engelen, M.; van Rijen, H.V.; Doevendans, P.A.; Crijns, H.J.; Ackerman, S.L.; Sluiter, W.; et al. EUK-8, a superoxide dismutase and catalase mimetic, reduces cardiac oxidative stress and ameliorates pressure overload-induced heart failure in the harlequin mouse mutant. J. Am. Coll. Cardiol. 2006, 48, 824–832. [Google Scholar] [CrossRef] [Green Version]
  204. Wang, Y.; Tabas, I. Emerging roles of mitochondria ROS in atherosclerotic lesions: Causation or association? J. Atheroscler. Thromb. 2014, 21, 381–390. [Google Scholar] [CrossRef] [Green Version]
  205. Escribano-Lopez, I.; Diaz-Morales, N.; Rovira-Llopis, S.; de Marañon, A.M.; Orden, S.; Alvarez, A.; Bañuls, C.; Rocha, M.; Murphy, M.P.; Hernandez-Mijares, A.; et al. The mitochondria-targeted antioxidant MitoQ modulates oxidative stress, inflammation and leukocyte-endothelium interactions in leukocytes isolated from type 2 diabetic patients. Redox Biol. 2016, 10, 200–205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Karnewar, S.; Pulipaka, S.; Katta, S.; Panuganti, D.; Neeli, P.K.; Thennati, R.; Jerald, M.K.; Kotamraju, S. Mitochondria-targeted esculetin mitigates atherosclerosis in the setting of aging via the modulation of SIRT1-mediated vascular cell senescence and mitochondrial function in Apoe−/− mice. Atherosclerosis 2022, 356, 28–40. [Google Scholar] [CrossRef] [PubMed]
  207. Vives-Bauza, C.; Zhou, C.; Huang, Y.; Cui, M.; De Vries, R.L.A.; Kim, J.; May, J.; Tocilescu, M.A.; Liu, W.; Ko, H.S.; et al. PINK1-dependent recruitment of Parkin to mitochondria in mitophagy. Proc. Natl. Acad. Sci. USA 2010, 107, 378–383. [Google Scholar] [CrossRef] [Green Version]
  208. Bravo-San Pedro, J.M.; Kroemer, G.; Galluzzi, L. Autophagy and Mitophagy in Cardiovascular Disease. Circ. Res. 2017, 120, 1812–1824. [Google Scholar] [CrossRef] [PubMed]
  209. Galluzzi, L.; Baehrecke, E.H.; Ballabio, A.; Boya, P.; Bravo-San Pedro, J.M.; Cecconi, F.; Choi, A.M.; Chu, C.T.; Codogno, P.; Colombo, M.I.; et al. Molecular definitions of autophagy and related processes. EMBO J. 2017, 36, 1811–1836. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Pankiv, S.; Clausen, T.H.; Lamark, T.; Brech, A.; Bruun, J.A.; Outzen, H.; Øvervatn, A.; Bjørkøy, G.; Johansen, T. p62/SQSTM1 binds directly to Atg8/LC3 to facilitate degradation of ubiquitinated protein aggregates by autophagy. J. Biol. Chem. 2007, 282, 24131–24145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  211. Bader, V.; Winklhofer, K.F. PINK1 and Parkin: Team players in stress-induced mitophagy. Biol. Chem. 2020, 401, 891–899. [Google Scholar] [CrossRef]
  212. Gustafsson, Å.B.; Dorn, G.W. Evolving and Expanding the Roles of Mitophagy as a Homeostatic and Pathogenic Process. Physiol. Rev. 2019, 99, 853–892. [Google Scholar] [CrossRef]
  213. Lampert, M.A.; Orogo, A.M.; Najor, R.H.; Hammerling, B.C.; Leon, L.J.; Wang, B.J.; Kim, T.; Sussman, M.A.; Gustafsson, Å.B. BNIP3L/NIX and FUNDC1-mediated mitophagy is required for mitochondrial network remodeling during cardiac progenitor cell differentiation. Autophagy 2019, 15, 1182–1198. [Google Scholar] [CrossRef] [PubMed]
  214. Marinković, M.; Šprung, M.; Novak, I. Dimerization of mitophagy receptor BNIP3L/NIX is essential for recruitment of autophagic machinery. Autophagy 2021, 17, 1232–1243. [Google Scholar] [CrossRef] [PubMed]
  215. Hamacher-Brady, A.; Brady, N.R.; Logue, S.E.; Sayen, M.R.; Jinno, M.; Kirshenbaum, L.A.; Gottlieb, R.A.; Gustafsson, Å.B. Response to myocardial ischemia/reperfusion injury involves Bnip3 and autophagy. Cell Death Differ. 2007, 14, 146–157. [Google Scholar] [CrossRef] [Green Version]
  216. Chen, M.; Chen, Z.; Wang, Y.; Tan, Z.; Zhu, C.; Li, Y.; Han, Z.; Chen, L.; Gao, R.; Liu, L.; et al. Mitophagy receptor FUNDC1 regulates mitochondrial dynamics and mitophagy. Autophagy 2016, 12, 689–702. [Google Scholar] [CrossRef] [Green Version]
  217. Fu, Z.J.; Wang, Z.Y.; Xu, L.; Chen, X.H.; Li, X.X.; Liao, W.T.; Ma, H.K.; Jiang, M.D.; Xu, T.T.; Xu, J.; et al. HIF-1α-BNIP3-mediated mitophagy in tubular cells protects against renal ischemia/reperfusion injury. Redox Biol. 2020, 36, 101671. [Google Scholar] [CrossRef]
  218. Xu, D.; Chen, P.; Wang, B.; Wang, Y.; Miao, N.; Yin, F.; Cheng, Q.; Zhou, Z.; Xie, H.; Zhou, L.; et al. NIX-mediated mitophagy protects against proteinuria-induced tubular cell apoptosis and renal injury. Am. J. Physiol. Ren. Physiol. 2019, 316, F382–F395. [Google Scholar] [CrossRef]
  219. Yao, M.; Liu, Y.; Sun, M.; Qin, S.; Xin, W.; Guan, X.; Zhang, B.; He, T.; Huang, Y. The molecular mechanisms and intervention strategies of mitophagy in cardiorenal syndrome. Front. Physiol. 2022, 13, 1008517. [Google Scholar] [CrossRef]
  220. Bhatia, D.; Chung, K.P.; Nakahira, K.; Patino, E.; Rice, M.C.; Torres, L.K.; Muthukumar, T.; Choi, A.M.K.; Akchurin, O.M.; Choi, M.E. Mitophagy-dependent macrophage reprogramming protects against kidney fibrosis. JCI Insight 2019, 4, 132826. [Google Scholar] [CrossRef] [Green Version]
  221. Duan, P.; Tan, J.; Miao, Y.; Zhang, Q. PINK1/Parkin-Mediated Mitophagy Plays a Protective Role in Albumin Overload-Induced Renal Tubular Cell Injury. Front. Biosci. 2022, 27, 184. [Google Scholar] [CrossRef] [PubMed]
  222. Tang, C.; Han, H.; Yan, M.; Zhu, S.; Liu, J.; Liu, Z.; He, L.; Tan, J.; Liu, Y.; Liu, H.; et al. PINK1-PRKN/PARK2 pathway of mitophagy is activated to protect against renal ischemia-reperfusion injury. Autophagy 2018, 14, 880–897. [Google Scholar] [CrossRef] [Green Version]
  223. Wang, Y.; Tang, C.; Cai, J.; Chen, G.; Zhang, D.; Zhang, Z.; Dong, Z. PINK1/Parkin-mediated mitophagy is activated in cisplatin nephrotoxicity to protect against kidney injury. Cell Death Dis. 2018, 9, 1113. [Google Scholar] [CrossRef] [Green Version]
  224. Zhao, C.; Chen, Z.; Xu, X.; An, X.; Duan, S.; Huang, Z.; Zhang, C.; Wu, L.; Zhang, B.; Zhang, A.; et al. Pink1/Parkin-mediated mitophagy play a protective role in cisplatin induced renal tubular epithelial cells injury. Exp. Cell Res. 2017, 350, 390–397. [Google Scholar] [CrossRef]
  225. Ishihara, M.; Urushido, M.; Hamada, K.; Matsumoto, T.; Shimamura, Y.; Ogata, K.; Inoue, K.; Taniguchi, Y.; Horino, T.; Fujieda, M.; et al. Sestrin-2 and BNIP3 regulate autophagy and mitophagy in renal tubular cells in acute kidney injury. Am. J. Physiol. Ren. Physiol. 2013, 305, F495–F509. [Google Scholar] [CrossRef]
  226. Wesselborg, S.; Stork, B. Autophagy signal transduction by ATG proteins: From hierarchies to networks. Cell. Mol. Life Sci. 2015, 72, 4721–4757. [Google Scholar] [CrossRef] [Green Version]
  227. Kawakami, T.; Gomez, I.G.; Ren, S.; Hudkins, K.; Roach, A.; Alpers, C.E.; Shankland, S.J.; D’Agati, V.D.; Duffield, J.S. Deficient autophagy results in mitochondrial dysfunction and FSGS. J. Am. Soc. Nephrol. 2015, 26, 1040–1052. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Cui, J.; Shi, S.; Sun, X.; Cai, G.; Cui, S.; Hong, Q.; Chen, X.; Bai, X.Y. Mitochondrial autophagy involving renal injury and aging is modulated by caloric intake in aged rat kidneys. PLoS ONE 2013, 8, e69720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Chen, K.; Dai, H.; Yuan, J.; Chen, J.; Lin, L.; Zhang, W.; Wang, L.; Zhang, J.; Li, K.; He, Y. Optineurin-mediated mitophagy protects renal tubular epithelial cells against accelerated senescence in diabetic nephropathy. Cell Death Dis. 2018, 9, 105. [Google Scholar] [CrossRef] [Green Version]
  230. Zheng, T.; Wang, H.Y.; Chen, Y.; Chen, X.; Wu, Z.L.; Hu, Q.Y.; Sun, H. Src Activation Aggravates Podocyte Injury in Diabetic Nephropathy via Suppression of FUNDC1-Mediated Mitophagy. Front. Pharmacol. 2022, 13, 897046. [Google Scholar] [CrossRef] [PubMed]
  231. Xiao, L.; Xu, X.; Zhang, F.; Wang, M.; Xu, Y.; Tang, D.; Wang, J.; Qin, Y.; Liu, Y.; Tang, C.; et al. The mitochondria-targeted antioxidant MitoQ ameliorated tubular injury mediated by mitophagy in diabetic kidney disease via Nrf2/PINK1. Redox Biol. 2017, 11, 297–311. [Google Scholar] [CrossRef]
  232. Han, P.; Yuan, C.; Wang, Y.; Wang, M.; Weng, W.; Zhan, H.; Yu, X.; Wang, T.; Li, Y.; Yi, W.; et al. Niclosamide ethanolamine protects kidney in adriamycin nephropathy by regulating mitochondrial redox balance. Am. J. Transl. Res. 2019, 11, 855–864. [Google Scholar]
  233. Han, P.; Zhan, H.; Shao, M.; Wang, W.; Song, G.; Yu, X.; Zhang, C.; Ge, N.; Yi, T.; Li, S.; et al. Niclosamide ethanolamine improves kidney injury in db/db mice. Diabetes Res. Clin. Pract. 2018, 144, 25–33. [Google Scholar] [CrossRef]
  234. Han, P.; Shao, M.; Guo, L.; Wang, W.; Song, G.; Yu, X.; Zhang, C.; Ge, N.; Yi, T.; Li, S.; et al. Niclosamide ethanolamine improves diabetes and diabetic kidney disease in mice. Am. J. Transl. Res. 2018, 10, 1071–1084. [Google Scholar]
  235. Zhang, L.X.; Zhao, H.J.; Sun, D.L.; Gao, S.L.; Zhang, H.M.; Ding, X.G. Niclosamide attenuates inflammatory cytokines via the autophagy pathway leading to improved outcomes in renal ischemia/reperfusion injury. Mol. Med. Rep. 2017, 16, 1810–1816. [Google Scholar] [CrossRef] [Green Version]
  236. Chang, X.; Zhen, X.; Liu, J.; Ren, X.; Hu, Z.; Zhou, Z.; Zhu, F.; Ding, K.; Nie, J. The antihelmenthic phosphate niclosamide impedes renal fibrosis by inhibiting homeodomain-interacting protein kinase 2 expression. Kidney Int. 2017, 92, 612–624. [Google Scholar] [CrossRef] [PubMed]
  237. Jiang, X.; Chen, X.; Hua, W.; He, J.; Liu, T.; Li, X.; Wan, J.; Gan, H.; Du, X. PINK1/Parkin mediated mitophagy ameliorates palmitic acid-induced apoptosis through reducing mitochondrial ROS production in podocytes. Biochem. Biophys. Res. Commun. 2020, 525, 954–961. [Google Scholar] [CrossRef] [PubMed]
  238. Kong, Z.L.; Che, K.; Hu, J.X.; Chen, Y.; Wang, Y.Y.; Wang, X.; Wen-Shan, L.; Wang, Y.G.; Chi, J.W. Orientin Protects Podocytes from High Glucose Induced Apoptosis through Mitophagy. Chem. Biodivers. 2020, 17, e1900647. [Google Scholar] [CrossRef] [PubMed]
  239. Glick, D.; Barth, S.; Macleod, K.F. Autophagy: Cellular and molecular mechanisms. J. Pathol. 2010, 221, 3–12. [Google Scholar] [CrossRef] [Green Version]
  240. Matsui, Y.; Takagi, H.; Qu, X.; Abdellatif, M.; Sakoda, H.; Asano, T.; Levine, B.; Sadoshima, J. Distinct roles of autophagy in the heart during ischemia and reperfusion: Roles of AMP-activated protein kinase and Beclin 1 in mediating autophagy. Circ. Res. 2007, 100, 914–922. [Google Scholar] [CrossRef] [Green Version]
  241. Tannous, P.; Zhu, H.; Johnstone, J.L.; Shelton, J.M.; Rajasekaran, N.S.; Benjamin, I.J.; Nguyen, L.; Gerard, R.D.; Levine, B.; Rothermel, B.A.; et al. Autophagy is an adaptive response in desmin-related cardiomyopathy. Proc. Natl. Acad. Sci. USA 2008, 105, 9745–9750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  242. Zhao, W.; Li, Y.; Jia, L.; Pan, L.; Li, H.; Du, J. Atg5 deficiency-mediated mitophagy aggravates cardiac inflammation and injury in response to angiotensin II. Free Radic. Biol. Med. 2014, 69, 108–115. [Google Scholar] [CrossRef] [PubMed]
  243. Sciarretta, S.; Zhai, P.; Shao, D.; Maejima, Y.; Robbins, J.; Volpe, M.; Condorelli, G.; Sadoshima, J. Rheb is a critical regulator of autophagy during myocardial ischemia: Pathophysiological implications in obesity and metabolic syndrome. Circulation 2012, 125, 1134–1146. [Google Scholar] [CrossRef] [Green Version]
  244. Dorn, G.W. Parkin-dependent mitophagy in the heart. J. Mol. Cell. Cardiol. 2016, 95, 42–49. [Google Scholar] [CrossRef] [Green Version]
  245. Lu, W.; Sun, J.; Yoon, J.S.; Zhang, Y.; Zheng, L.; Murphy, E.; Mattson, M.P.; Lenardo, M.J. Mitochondrial Protein PGAM5 Regulates Mitophagic Protection against Cell Necroptosis. PLoS ONE 2016, 11, 147792. [Google Scholar] [CrossRef] [Green Version]
  246. Andres, A.M.; Hernandez, G.; Lee, P.; Huang, C.; Ratliff, E.P.; Sin, J.; Thornton, C.A.; Damasco, M.V.; Gottlieb, R.A. Mitophagy is required for acute cardioprotection by simvastatin. Antioxid. Redox Signal. 2014, 21, 1960–1973. [Google Scholar] [CrossRef]
  247. Kubli, D.A.; Zhang, X.; Lee, Y.; Hanna, R.A.; Quinsay, M.N.; Nguyen, C.K.; Jimenez, R.; Petrosyans, S.; Murphy, A.N.; Gustafsson, Å.B. Parkin protein deficiency exacerbates cardiac injury and reduces survival following myocardial infarction. J. Biol. Chem. 2013, 288, 915–926. [Google Scholar] [CrossRef] [Green Version]
  248. Diwan, A.; Krenz, M.; Syed, F.M.; Wansapura, J.; Ren, X.; Koesters, A.G.; Li, H.; Kirshenbaum, L.A.; Hahn, H.S.; Robbins, J.; et al. Inhibition of ischemic cardiomyocyte apoptosis through targeted ablation of Bnip3 restrains postinfarction remodeling in mice. J. Clin. Investig. 2007, 117, 2825–2833. [Google Scholar] [CrossRef] [Green Version]
  249. Zhang, W.; Siraj, S.; Zhang, R.; Chen, Q. Mitophagy receptor FUNDC1 regulates mitochondrial homeostasis and protects the heart from I/R injury. Autophagy 2017, 13, 1080–1081. [Google Scholar] [CrossRef] [PubMed]
  250. Wu, X.; He, L.; Chen, F.; He, X.; Cai, Y.; Zhang, G.; Yi, Q.; He, M.; Luo, J. Impaired autophagy contributes to adverse cardiac remodeling in acute myocardial infarction. PLoS ONE 2014, 9, 0112891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  251. Kanamori, H.; Takemura, G.; Goto, K.; Maruyama, R.; Ono, K.; Nagao, K.; Tsujimoto, A.; Ogino, A.; Takeyama, T.; Kawaguchi, T.; et al. Autophagy limits acute myocardial infarction induced by permanent coronary artery occlusion. Am. J. Physiol. Heart Circ. Physiol. 2011, 300, H2261–H2271. [Google Scholar] [CrossRef] [Green Version]
  252. Santovito, D.; Egea, V.; Bidzhekov, K.; Natarelli, L.; Mourão, A.; Blanchet, X.; Wichapong, K.; Aslani, M.; Brunßen, C.; Horckmans, M.; et al. Noncanonical inhibition of caspase-3 by a nuclear microRNA confers endothelial protection by autophagy in atherosclerosis. Sci. Transl. Med. 2020, 12, eaaz2294. [Google Scholar] [CrossRef] [PubMed]
  253. Li, P.; Bai, Y.; Zhao, X.; Tian, T.; Tang, L.; Ru, J.; An, Y.; Wang, J. NR4A1 contributes to high-fat associated endothelial dysfunction by promoting CaMKII-Parkin-mitophagy pathways. Cell Stress Chaperones 2018, 23, 749–761. [Google Scholar] [CrossRef]
  254. He, L.; Zhou, Q.; Huang, Z.; Xu, J.; Zhou, H.; Lv, D.; Lu, L.; Huang, S.; Tang, M.; Zhong, J.; et al. PINK1/Parkin-mediated mitophagy promotes apelin-13-induced vascular smooth muscle cell proliferation by AMPKα and exacerbates atherosclerotic lesions. J. Cell. Physiol. 2019, 234, 8668–8682. [Google Scholar] [CrossRef]
  255. Zhang, X.; Sergin, I.; Evans, T.D.; Jeong, S.J.; Rodriguez-Velez, A.; Kapoor, D.; Chen, S.; Song, E.; Holloway, K.B.; Crowley, J.R.; et al. High-protein diets increase cardiovascular risk by activating macrophage mTOR to suppress mitophagy. Nat. Metab. 2020, 2, 110–125. [Google Scholar] [CrossRef] [PubMed]
  256. Zhang, W.; Ma, Q.; Siraj, S.; Ney, P.A.; Liu, J.; Liao, X.; Yuan, Y.; Li, W.; Liu, L.; Chen, Q. Nix-mediated mitophagy regulates platelet activation and life span. Blood Adv. 2019, 3, 2342–2354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Tyrrell, D.J.; Blin, M.G.; Song, J.; Wood, S.C.; Zhang, M.; Beard, D.A.; Goldstein, D.R. Age-Associated Mitochondrial Dysfunction Accelerates Atherogenesis. Circ. Res. 2020, 126, 298–314. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Therapeutic approaches and key targets in mitochondrial biogenesis in the cardio-renal axis. Dysfunction of mitochondrial biogenesis transcriptional activators such as the peroxisome proliferator-activated receptor γ-coactivator-1 family (PGC-1α/PGC-1β/PRC) and of transcription factor effectors (PPAR family, ERR family, NRF-1, and TFAM) that also regulate mitochondrial biogenesis may trigger deleterious responses in the kidney and cardiovascular system. (PGC-1α: Peroxisome proliferator-activated receptor-gamma coactivator 1-α/PGC-1β: Peroxisome proliferator-activated receptor-gamma coactivator 1-β/PRC: PGC-1-related coactivator/PPAR: Peroxisome proliferator-activated receptor/ERR: estrogen-related receptors/NRF-1: Nuclear Respiratory Factor 1/TFAM: mitochondrial transcription factor A/TWEAK: Tumor necrosis factor-like weak inducer of apoptosis/NAM: Nicotinamide/AMPK: AMP-activated protein kinase/PDE: phosphodiesterase/TNF-α: Tumor necrosis factor α/VSMCs: Vascular smooth muscle cells/AKI: Acute kidney injury/CKD: Chronic kidney disease).
Figure 1. Therapeutic approaches and key targets in mitochondrial biogenesis in the cardio-renal axis. Dysfunction of mitochondrial biogenesis transcriptional activators such as the peroxisome proliferator-activated receptor γ-coactivator-1 family (PGC-1α/PGC-1β/PRC) and of transcription factor effectors (PPAR family, ERR family, NRF-1, and TFAM) that also regulate mitochondrial biogenesis may trigger deleterious responses in the kidney and cardiovascular system. (PGC-1α: Peroxisome proliferator-activated receptor-gamma coactivator 1-α/PGC-1β: Peroxisome proliferator-activated receptor-gamma coactivator 1-β/PRC: PGC-1-related coactivator/PPAR: Peroxisome proliferator-activated receptor/ERR: estrogen-related receptors/NRF-1: Nuclear Respiratory Factor 1/TFAM: mitochondrial transcription factor A/TWEAK: Tumor necrosis factor-like weak inducer of apoptosis/NAM: Nicotinamide/AMPK: AMP-activated protein kinase/PDE: phosphodiesterase/TNF-α: Tumor necrosis factor α/VSMCs: Vascular smooth muscle cells/AKI: Acute kidney injury/CKD: Chronic kidney disease).
Ijms 24 08209 g001
Figure 2. Therapeutic approaches and key mitochondrial dynamics’ therapeutic targets in the cardio-renal axis. Modulators of mitochondrial fusion (MFN1/MFN2/OPA1) and fission (DRP1/FIS1) regulate mitochondrial dynamics, and defects on them may trigger deleterious effects in CKD and CVD. (FIS1: Mitochondrial fission 1 protein/DRP1: Dynamin-related protein 1/MFN1: Mitofusin-1/MNF2: Mitofusin-1/OPA1: protein optic atrophy 1/CVD: Cardiovascular disease/AKI: Acute kidney injury/CKD: Chronic kidney disease).
Figure 2. Therapeutic approaches and key mitochondrial dynamics’ therapeutic targets in the cardio-renal axis. Modulators of mitochondrial fusion (MFN1/MFN2/OPA1) and fission (DRP1/FIS1) regulate mitochondrial dynamics, and defects on them may trigger deleterious effects in CKD and CVD. (FIS1: Mitochondrial fission 1 protein/DRP1: Dynamin-related protein 1/MFN1: Mitofusin-1/MNF2: Mitofusin-1/OPA1: protein optic atrophy 1/CVD: Cardiovascular disease/AKI: Acute kidney injury/CKD: Chronic kidney disease).
Ijms 24 08209 g002
Figure 3. Oxidative stress is intrinsically linked to the physiopathology of cardio-renal syndrome. (1) ROS generated by damaged mitochondria induce cell death, inflammation, and fibrosis of renal and cardiovascular tissues. In turn, cellular stress may stimulate ROS production, establishing a vicious cycle between mitochondrial dysfunction and cell injury. (2) ROS production is concentrated in the inner mitochondrial membrane where the ETC is located. (3) Several electrons donated by the coenzymes NADH and FADH2 are transferred between the different components of the ETC (from complex I to V) with the transport of protons across the inner membrane, creating the electrochemical gradient that generates ATP. (4) The generation of ROS is characterized by the production of superoxide anion (O2) by the transfer of electrons to O2. O2 is then converted to hydrogen peroxide (H2O2) by the enzyme SOD in mitochondria, and H2O2 is converted to water by GPx or CAT. (5) Excess mitochondrial ROS production can ultimately oxidize mtDNA, proteins, and lipids. Interfering with electron transport by targeting the enzymatic complexes of the ETC and inhibiting cytoplasmic enzymes that lead to excessive free radical production has proven beneficial in limiting the pathological mechanisms triggered by damaged mitochondria. (ETC: electron transport chain/OXPHOS: Oxidative phosphorylation/NADH: nicotinamide adenine dinucleotide/FADH2: Flavin Adenine Dinucleotide (FAD) Reduced Form/ATP: Adenosine Tri-Phosphat/SOD: superoxide dismutase/GPx: glutathione peroxidase/ADP: Adenosine diphosphate/mCAT: mitochondrial catalase/ROS: Reactive oxygen species//mt DNA: mitochondrial DNA/mtOS: mitochondrial oxidation/CVD: Cardiovascular disease/AKI: Acute kidney injury/CKD: Chronic kidney disease/ΔΨm: mitochondrial membrane potential).
Figure 3. Oxidative stress is intrinsically linked to the physiopathology of cardio-renal syndrome. (1) ROS generated by damaged mitochondria induce cell death, inflammation, and fibrosis of renal and cardiovascular tissues. In turn, cellular stress may stimulate ROS production, establishing a vicious cycle between mitochondrial dysfunction and cell injury. (2) ROS production is concentrated in the inner mitochondrial membrane where the ETC is located. (3) Several electrons donated by the coenzymes NADH and FADH2 are transferred between the different components of the ETC (from complex I to V) with the transport of protons across the inner membrane, creating the electrochemical gradient that generates ATP. (4) The generation of ROS is characterized by the production of superoxide anion (O2) by the transfer of electrons to O2. O2 is then converted to hydrogen peroxide (H2O2) by the enzyme SOD in mitochondria, and H2O2 is converted to water by GPx or CAT. (5) Excess mitochondrial ROS production can ultimately oxidize mtDNA, proteins, and lipids. Interfering with electron transport by targeting the enzymatic complexes of the ETC and inhibiting cytoplasmic enzymes that lead to excessive free radical production has proven beneficial in limiting the pathological mechanisms triggered by damaged mitochondria. (ETC: electron transport chain/OXPHOS: Oxidative phosphorylation/NADH: nicotinamide adenine dinucleotide/FADH2: Flavin Adenine Dinucleotide (FAD) Reduced Form/ATP: Adenosine Tri-Phosphat/SOD: superoxide dismutase/GPx: glutathione peroxidase/ADP: Adenosine diphosphate/mCAT: mitochondrial catalase/ROS: Reactive oxygen species//mt DNA: mitochondrial DNA/mtOS: mitochondrial oxidation/CVD: Cardiovascular disease/AKI: Acute kidney injury/CKD: Chronic kidney disease/ΔΨm: mitochondrial membrane potential).
Ijms 24 08209 g003
Figure 4. Mitophagy and cardio-renal syndrome. (1) Under physiological conditions, PINK1 is cleaved by proteases in the mitochondria. (2) Insults that induce loss of mitochondrial membrane potential trigger PINK1 stabilization and activation in the outer mitochondrial membrane. (3) Activated PINK1 phosphorylates ubiquitin molecules and activates Parkin. Autophagy adaptors subsequently bind to the poly-ubiquitin chains and to LC3 inducing the phagophore formation near mitochondria. (4) and (5) Other mitophagy mediators in the outer mitochondrial membrane include BNIP3, NIX, and FUNDC1 that directly bind to LC3, triggering autophagosome engulfment. Dysfunction of these systems may contribute to CVD and kidney disease and may be targeted therapeutically by autophagy inducers. (PINK1:PTEN Induced Kinase 1/BNIP3: BCL2/adenovirus E1B 19 kDa protein-interacting protein 3/NIX: BNIP3-like protein/FUNDC1: FUN14 domain containing 1/LC3/PARKIN: Parkin RBR E3 Ubiquitin Protein Ligase/3-MA: 3-Methyladenine/BafA1: Bafilomycin A1/ATG5-8: Autophagy-related-5–8/VSMCs: Vascular smooth muscle cells/MitoQ: Mitoquinone mesylate/CVD: Cardiovascular disease/AKI: Acute kidney injury/CKD: Chronic kidney disease).
Figure 4. Mitophagy and cardio-renal syndrome. (1) Under physiological conditions, PINK1 is cleaved by proteases in the mitochondria. (2) Insults that induce loss of mitochondrial membrane potential trigger PINK1 stabilization and activation in the outer mitochondrial membrane. (3) Activated PINK1 phosphorylates ubiquitin molecules and activates Parkin. Autophagy adaptors subsequently bind to the poly-ubiquitin chains and to LC3 inducing the phagophore formation near mitochondria. (4) and (5) Other mitophagy mediators in the outer mitochondrial membrane include BNIP3, NIX, and FUNDC1 that directly bind to LC3, triggering autophagosome engulfment. Dysfunction of these systems may contribute to CVD and kidney disease and may be targeted therapeutically by autophagy inducers. (PINK1:PTEN Induced Kinase 1/BNIP3: BCL2/adenovirus E1B 19 kDa protein-interacting protein 3/NIX: BNIP3-like protein/FUNDC1: FUN14 domain containing 1/LC3/PARKIN: Parkin RBR E3 Ubiquitin Protein Ligase/3-MA: 3-Methyladenine/BafA1: Bafilomycin A1/ATG5-8: Autophagy-related-5–8/VSMCs: Vascular smooth muscle cells/MitoQ: Mitoquinone mesylate/CVD: Cardiovascular disease/AKI: Acute kidney injury/CKD: Chronic kidney disease).
Ijms 24 08209 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mendez-Barbero, N.; Oller, J.; Sanz, A.B.; Ramos, A.M.; Ortiz, A.; Ruiz-Ortega, M.; Rayego-Mateos, S. Mitochondrial Dysfunction in the Cardio-Renal Axis. Int. J. Mol. Sci. 2023, 24, 8209. https://doi.org/10.3390/ijms24098209

AMA Style

Mendez-Barbero N, Oller J, Sanz AB, Ramos AM, Ortiz A, Ruiz-Ortega M, Rayego-Mateos S. Mitochondrial Dysfunction in the Cardio-Renal Axis. International Journal of Molecular Sciences. 2023; 24(9):8209. https://doi.org/10.3390/ijms24098209

Chicago/Turabian Style

Mendez-Barbero, Nerea, Jorge Oller, Ana B. Sanz, Adrian M. Ramos, Alberto Ortiz, Marta Ruiz-Ortega, and Sandra Rayego-Mateos. 2023. "Mitochondrial Dysfunction in the Cardio-Renal Axis" International Journal of Molecular Sciences 24, no. 9: 8209. https://doi.org/10.3390/ijms24098209

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