Therapeutic Strategies Targeting Mitochondrial Metabolism in Cardiovascular Disease: History
Please note this is an old version of this entry, which may differ significantly from the current revision.
Contributor: , , , ,

Cardiovascular disease (CVD) is a group of systemic disorders threatening human health with complex pathogenesis, among which mitochondrial energy metabolism reprogramming has a critical role. Mitochondria are cell organelles that fuel the energy essential for biochemical reactions and maintain normal physiological functions of the body. Mitochondrial metabolic disorders are extensively involved in the progression of CVD, especially for energy-demanding organs such as the heart. Therefore, elucidating the role of mitochondrial metabolism in the progression of CVD is of great significance to further understand the pathogenesis of CVD and explore preventive and therapeutic methods.

  • cardiovascular disease
  • mitochondrial metabolism
  • mitochondrial calcium

1. Introduction

Cardiovascular disease (CVD) remains the leading cause of death and disease burden worldwide [1]. Energy-demanding organs such as the heart are affected by mitochondrial function, and the relationship between mitochondrial metabolism and CVD has been widely proved [2][3][4][5][6][7]. Mitochondrial metabolism is a highly complex energy-releasing process involving a series of enzymatic reactions, during which sugars, fats, and proteins are oxidized through the tricarboxylic acid cycle (TCA) and oxidative phosphorylation (OXPHOS). Mitochondrial dysfunction is a symbol of the exacerbated intracellular environment leading to diseases such as CVD [2][3][4][5][6][7]. Regulating mitochondrial metabolism is a promising therapeutic strategy for CVD.

2. Mitochondrial Metabolism Dysfunction in CVD

Mitochondrial metabolism is the main source of energy supply for most cells, especially for energy-consuming cells, such as cardiomyocytes. The glucose OXPHOS in the mitochondria is the major metabolism pathway in most cells. OXPHOS is, indeed, a more efficient means of generating ATP than glycolysis [8]. Mitochondrial metabolic dysfunction has been widely reported in CVD, characterized by metabolic reprogramming [9]. In pulmonary arterial hypertension (PAH), a significant metabolic reprogramming is observed, that is, in pulmonary artery smooth muscle cells (PASMCs), the level of OXPHOS is significantly reduced, and the levels of glycolysis and fatty acid oxidation (FAO) are significantly increased [10][11]. Similar metabolic reprogramming from glucose OXPHOS to glycolysis occurs in aortic smooth muscle cells (SMCs) in the development of aneurysms and atherosclerotic diseases [5][12]. The major metabolism pathway in cardiomyocytes is FAO. During the progression of heart failure, metabolism shifts from majorly relying on FAO to OXPHOS in the early stage then leans to glycolysis in the late stage [13].

3. Major Factors in Mitochondrial Metabolism

3.1. Mitochondrial DNA

Mitochondria are the only semi-autonomous genetic organelles in mammalian cells with their own transcriptional and translational system independent of the nucleus. Mitochondria autonomously synthesize a variety of key proteins and enzymes by mitochondrial DNA (mtDNA). mtDNA is mainly distributed in the mitochondrial matrix and inner membrane and contains a double-stranded circular DNA with a molecular weight of 16.5 kb, 16,569 nucleotide pairs, coding sequences of 37 genes, 2 ribosomal RNAs, 22 transporter RNAs, and 13 genes encoding important component proteins that are closely related to mitochondrial OXPHOS [14][15]. mtDNA plays an extremely important role in providing the template for translating enzymes that maintain the normal mitochondrial OXPHOS process and cell function. The enzyme includes 7 mitochondrial oxidative respiratory chain NADH dehydrogenase subunits, 1 cytochrome B, 3 cytochrome C oxidase subunits, and 2 ATP synthetases [16]. Due to the absence of introns and histones in mtDNA, and the lack of a similar DNA repair system in the nucleus, mtDNA is more susceptible to oxidative damage by reactive oxygen species (ROS) and reactive nitrogen species (RNS) [17].

3.2. Mitochondrial Dynamics

Mitochondrial dynamics refers to the dynamic cycle of biogenesis, fusion, fission, and degradation that mitochondria continuously undergo to maintain their integrity, distribution, and morphology [18]. These processes reflect the early adaptation of mitochondria in response to external stimuli, maintaining overall mitochondrial function by removing irreversibly damaged parts and preserving the functional parts. In different tissues and cells, mitochondria morphology and network structure vary. In cardiomyocytes, mitochondria are numerous, covered with ridges, and spherical in shape, while in lymphocytes, the mitochondria are few in the count, tube-shaped. This also shows that the correlation of different morphological features of mitochondria with the energy requirements of tissue cells. Increasing evidences show that the disturbance of mitochondrial dynamics is extensively involved in the progression of CVD, including heart failure [19][20], myocardial ischemia/reperfusion (I/R) injury [21], abdominal aortic aneurysm (AAA) [22] and PAH [23][24].

3.2.1. Mitochondrial Fusion

Mitochondrial fusion is an important process that enables the exchange of DNA, proteins, lipids, and metabolites between different mitochondria [25]. Mitochondrial fusion consists of two processes, outer and inner membrane fusion, regulated by the nuclear-encoded dynamin-related GTPases protein family membrane fusion proteins 1 (Mfn1), membrane fusion proteins 2 (Mfn2), and optic atrophy protein 1 (OPA1). The outer membrane fusion is dependent on Mfn1 and Mfn2, while OPA1, located in the inner mitochondrial membrane, is involved in the inner mitochondrial membrane fusion. Numerous studies have shown that mitochondrial fusion is involved in mitochondrial metabolism [26]. Mfn2 can interact with pyruvate kinase isozyme type M2 (PKM2) to promote mitochondrial fusion, facilitate OXPHOS, and inhibit glycolysis [27]. In both yeast and mammalian cells, blocking mitochondrial fusion can lead to mitochondrial dysfunction and inadequate ATP supply [28]. Loss-of-function mutations of Mfn1 and Mfn2 cause: (1) decreased cellular glucose oxidation level, (2) depolarization of the mitochondrial membrane potential, (3) inhibition of the mitochondrial TCA and oxidative respiration, (4) conversion of cellular energy metabolism from OXPHOS to glycolysis, and (5) severe cellular functional defects. Whereas stress-induced mitochondrial fusion restores mitochondrial ATP production [29]. When mitochondrial OXPHOS is blocked, it causes an increase in mitochondrial ROS (mtROS), which triggers oxidative damage to mitochondria and increases mtDNA mutations.

3.2.2. Mitochondrial Fission

Mitochondrial fission is mainly mediated by mitochondrial fission protein 1 (Fis1) and dynamin-related protein 1 (Drp1) in the cytoplasm. Under physiological conditions, mitochondrial division can serve the purpose of isolating irreversibly damaged DNA or toxic material from the mitochondria, and defective mitochondria are removed via the mitophagy pathway when the mitochondrial membrane potential and pH gradient fail to meet the requirements for fusion. Thus, moderate mitochondrial division is beneficial to maintain normal cellular energy metabolism. However, excessive mitochondrial fission inhibits the respiratory chain, represses ATP production, and leads to cellular dysfunction [30][31].
There is an imbalance in [Ca2+]m homeostasis is an important cause of CVD [32]. For example, heart-specific knockdown of VDAC2 has been found to cause [Ca2+]m imbalance and dilated cardiomyopathy [33]. Similarly, a lack of MCU has been found to promote PASMC proliferation and apoptosis resistance in human and experimental PAH models, and restoring MCU expression can alleviate the symptoms of experimental PAH [34]. Therefore, the maintenance of [Ca2+]m homeostasis is essential to maintaining the normal energy metabolism of mitochondria and the function of the body. [Ca2+]m imbalance can be caused by [Ca2+]m overload. Numerous studies have shown that when [Ca2+]m is abnormally increased, may cause [Ca2+]m overload, resulting in oxidative stress, mitochondrial structural damage, and dysfunction of mitochondria, which in turn reduces ATP production and triggers cell apoptosis [2][35][36][37][38]. Nevertheless, [Ca2+]m itself is a key factor in regulating mitochondrial ATP production by participating in the following processes. Ca2+ can activate pyruvate dehydrogenase complex (PDC), a complex enzyme composed of multiple subunits [39], isocitrate dehydrogenase (IDH), a key TCA rate-limiting enzyme [40], and 2-oxoglutarate dehydrogenase complex (OGDC) in the mitochondrial matrix [41][42][43]. When glucose is converted to pyruvate in the cytosol, it is transported to the mitochondria and catalyzed by PDC to acetyl coenzyme A (acetyl-CoA). Acetyl-CoA undergoes the TCA producing carbon dioxide, GTP, and NADH. Ca2+ can enhance PDC activity by binding to pyruvate dehydrogenase (PDH) phosphatase subunit 1 (PDP1), which accelerates the conversion of pyruvate to acetyl-CoA [44][45] and promote the ATP production through dephosphorylation of the PDH subunit [46]. Meanwhile, Ca2+ can accelerate the TCA by enhancing the activity of IDH [47] and OGDC [48].

3.4. ROS

ROS refer to a class of free radicals composed of oxygen molecules with an odd number of electrons, characterized by a short half-life and high activity. Mitochondria are the main source of ROS when electrons leak during transfer and are captured by oxygen molecules. During OXPHOS, mitochondria transfer electrons from electron donors to electron acceptors through the electron transfer chain, catalyze the generation of ATP and reduce oxygen molecules to water. Some studies have speculated that about 0.2–2% of these consumed oxygen molecules are converted to ROS [49]. In physiological conditions, low levels of ROS can act as signaling molecules that participate in mitochondrial metabolism [50][51][52][53][54], and promote adaptive upregulation of antioxidant enzymes to maintain the body’s health [55]. ROS could be rapidly scavenged by antioxidant enzymes, such as superoxide dismutase 2 (SOD2), glutathione reductase, and catalase. When ROS exceed the clearance capacity of the antioxidant system, excessive ROS may damage proteins, lipids, and mtDNA in mitochondria causing oxidative stress and diseases [56], such as CVD [57]. Therefore, to avoid cellular dysfunction, redox homeostasis must be strictly controlled in case of excessive ROS accumulation [58].

4. Mitochondrial Metabolism Disorder and CVD

4.1. PAH

PAH is a heterogeneous and fatal disease of the pulmonary vasculature, clinically defined as resting mean pulmonary artery pressure >20 mmHg, and normal left atrial pressure and pulmonary vascular resistance ≥3 Wood units [59]. PAH is characterized by occlusive remodeling and increased resistance caused by persistent pulmonary vasoconstriction, leading to right ventricular hypertrophy and heart failure. Excessive proliferation and apoptosis resistance of PASMCs and pulmonary artery endothelial cells (PAECs) are the main causes of pulmonary vascular remodeling.
The association of metabolic reprogramming and PAH has become a consensus over the past decades [60]. In PAH, a shift of energy metabolism from OXPHOS of glucose to cytoplasmic glycolysis in PASMCs and PAECs has been widely demonstrated in human lung-derived cell cultures and multiple mouse models of PAH, and the prevailing view is that the driving force for this metabolic shift stems from the pathological accumulation of hypoxia-inducible factor 1α (HIF-1α) [61]. HIF-1α is a hypoxia-inducible protein that significantly promotes glycolytic metabolism and is a core regulator of the body’s adaptive repair of the intracellular oxygen environment, playing an important role in a variety of physiological functions including cell proliferation, pro-survival, angiogenesis and metabolism. In the development of PAH, HIF-1α accumulation is accompanied by an increase in glycolysis and glycolysis-related enzyme activities [10]. In contrast, mitochondrial glucose OXPHOS decreases significantly in the pulmonary artery. However, it is noteworthy that the levels of enzymes related to fatty acid metabolism and FAO in mitochondria of PASMCs and PAECs are significantly increased in a variety of PAH models [62][63][64].

4.2. Aortic Aneurysm

Aortic aneurysm is an aortic dilatation disease caused by multiple factors, characterized by irreversible dilatation of the aortic diameter beyond 50% of the normal diameter or the abdominal aortic diameter ≥3 cm. Risk factors for aneurysms include smoking, age, and genetics. When an aortic aneurysm occurs, the risk of aneurysm rupture gradually increases parallel with the continued expansion of the aorta and the enlargement of the aneurysm. Once the aneurysm ruptures, the chance of fatality will be greater than 80% [65]. Moreover, due to the insidious nature of aneurysm development, patients may not show any clinical symptoms before the aneurysm rupture. Therefore, aneurysm rupture is the greatest risk factor for death in patients with an aneurysm. Because of the lack of a specific drug and the short time from onset to death, early screening is an important clinical precaution method, for aneurysms larger than 5.5 cm in diameter, the main treatment is endovascular aortic repair and surgery. It is urgent to determine the pathogenesis of aneurysms and develop targeted drugs. The main features of aneurysms are excessive proliferation and migration of aortic SMCs, increased secretion of extracellular matrix and degradation of elastic fiber, which cause vascular remodeling. Inflammation and redox imbalance are considered important drivers of aneurysm development. Chronic vascular inflammation can recruit immune cells, amplify the inflammatory response, cause SMCs dysfunction and apoptosis, and eventually lead to aortic smooth muscle tissue damage, causing an aneurysm [66].

4.3. Atherosclerosis

Atherosclerosis, a vascular disease characterized by lipid deposition and inflammation in the arterial wall, is the pathological cause of coronary heart disease and peripheral vascular diseases imposing a serious risk to health. Vascular endothelial cells (ECs), macrophages, and SMCs are the main cell types chronologically involved in the development of atherosclerosis. As initiates with hypercholesterolemia-induced endothelial dysfunction [67], damaged ECs trigger inflammation to recruit monocytes. Monocytes are induced to M1 macrophages, which accelerate the local inflammatory reaction and promote abnormal proliferation of SMCs leading to formation of atherosclerosis.
Glycolysis is the main form of energy metabolism in ECs and M1 macrophages [68][69]. Studies have found that the expression of glycolysis-related enzymes and glycolysis in ECs are significantly down-regulated in atherosclerosis [70]. In contrast, in atherosclerosis, inflammatory factors can promote the metabolic reprogramming of M1 macrophages from myelocytomatosis viral oncogene-dependent to HIF-1α-dependent, which enhances glycolysis [71]. While under hypoxia, either knockdown of HIF-1α or inhibition of the expression of glycolytic related protein 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase 3 (PFKFB3) can significantly hamper the glycolysis in macrophages, reduce the expression of tumor necrosis factor-alpha (TNF-α), and promote macrophage apoptosis [72].

4.4. Heart Failure

Cardiomyocytes contains large numbers of mitochondria as heart is one of the most energy-consuming organs of the body. Thus, disturbances in mitochondrial metabolism can lead to an inadequate energy supply to the heart and cause cardiac dysfunction. The reprogramming of cardiac energy metabolism during the development of heart failure has been widely demonstrated [13][73][74]. Under physiological conditions, about 40–60% of the energy required for myocardial activity derives from mitochondrial FAO, and the rest mainly comes from the oxidation of sugars, ketones and amino acids [75].

5. Mitochondria-Targeted Therapeutic Agents and Strategies in CVD

Current strategies for mitochondrially targeted therapy mainly focus on factors of mitochondrial damage, such as decreased ATP/ADP ratio, signal pathway changes caused by insufficient NAD+, ROS-induced metabolic disorders, and [Ca2+]m disorders caused by abnormal calcium channels which lead to strategies for mitochondria-targeted drug delivery: (1) cationic carriers designed using mitochondrial membrane potential coupled with drugs, such as mitoquinone (MitoQ) coupled with lipophilic triphenylphosphonium (TPP+), (2) mitochondrially targeted signal peptides as carriers for drug delivery, such as Szeto-Schiller peptide 31 (SS-31), (3) liposomes as carriers, such as mitochondria-targeted dendritic lipopeptide liposomal delivery platform [76], (4) nanoparticles coated with drugs for delivery, such as mitochondria-targeted nanoparticles (CsA@PLGA-PEG-SS31) [77], (5) physical invasion methods, such as microinjection of mitochondria directly into cells [78], (6) gene therapy technology, such as mitoTALENs technology [79]

5.1. Potential Mitochondrial Targets for CVD

5.1.1. MPTP

MPTP, a complex assembled by a group of proteins, is a non-specific channel located between the inner and outer membranes of mitochondria. Numerous studies have shown that MPTP plays important roles in mitochondria-induced apoptosis [80][81]. Under stress conditions, mitochondrial oxidative stress can lead to [Ca2+]m overload by increasing [Ca2+]m uptake, causing the opening of MPTP, which in turn leads to a decrease in mitochondrial membrane potential and release of cytochrome C, triggering of mitochondrial apoptosis signal to induce cell apoptosis [82][83].

5.1.2. Sirtuin 3 (Sirt3)

Acetylation of mitochondrial proteins is involved in the pathogenesis of CVD [84]. Sirt3 is a member of the histone deacetylase sirtuin family localized in mitochondria. Proteomic analysis reveals that most of the mitochondrial proteins can be modified by acetylation, and Sirt3 plays an important role in the deacetylation of these proteins [85].

5.1.3. [Ca2+]m

[Ca2+]m homeostasis is vital for mitochondrial metabolism and CVD [33][34]. In recent years, methods of restoring [Ca2+]m imbalance, namely modulating mitochondrial calcium uptakes/release channels such as VDACs, MCU or NLX, have attracted a lot of attentions. It is considered that [Ca2+]m overload caused by MCU complex activation is the underlying mechanism of cardiac I/R injury.

5.1.4. Mitochondrial Dynamics

Mitochondrial dynamics is mainly characterized by a dynamic balance between mitochondrial fission and fusion. Mitochondrial fission and fusion have an important role in the mitochondrial material exchange, repair, and removal of damaged mitochondria, and are an important means for the body to maintain mitochondrial health. Extreme states of excessive fission or fusion are detrimental to health [86]. Under physiological or pathological conditions, mitochondrial fission and fusion form new homeostasis to meet the energy or physiological needs of the cell.

5.2. Mitochondria-Targeted Agents for CVD

5.2.1. CoQ10 and MitoQ

CoQ10 is one of the electron carriers in the mitochondrial ETC. It is involved in mediating the electron transfer between mitochondrial OXPHOS complex I/II and complex III and plays an important role in maintaining mitochondrial aerobic respiration and ATP production. CoQ10 is widely used in the treatment of metabolism-related diseases because of its powerful antioxidant protective effect, which protects mitochondria from oxidative damage to maintain normal mitochondrial metabolic function. Due to the double membrane structure of mitochondria, CoQ10 cannot effectively enter mitochondria, resulting in low bioavailability. Scientists synthesize MitoQ with mitochondria-targeted function by covalently coupling the benzoquinone part of CoQ10 and TPP+ through the deca-carbon aliphatic chain by taking advantage of the mitochondrial membrane potential targeting property of TPP+. MitoQ has been widely used in the treatment of various metabolic syndromes [87].

5.2.2. Melatonin

Melatonin is a neuroendocrine hormone produced by the pineal gland of mammals and its physiological functions include neuroregulation, sleep promotion, antioxidant, anti-inflammation, anti-aging, immune regulation, endocrine regulation and cell growth promotion [88]. The level of melatonin varies greatly among different populations and decreases gradually with aging. Melatonin has amphipathic properties enabling itself penetrating through a variety of cell membranes. Melatonin can quickly cross the cellular and mitochondrial membranes and accumulate in mitochondria [89]. Due to its powerful antioxidant and anti-inflammatory properties, melatonin has been widely used in therapeutic studies of CVD. Encouragingly, most studies have found a protective effect of melatonin in heart failure [90], cardiac I/R injury [91], aneurysm [92][93], atherosclerosis [94], and PAH [95][96].

5.2.3. SS-31

SS-31 is widely used in CVD treatment. In the aging heart, SS-31 reduces oxidative stress and mitochondrial electron leakage, inhibits the opening of MPTP, improves cardiac energy metabolism, restores mitochondrial activity, promotes ATP generation, and alleviates cardiac diastolic function [97][98][99]. Cardiac proteomics studies have found that SS-31 ameliorates aging-induced changes in the post-translational modification profile of cardiac proteins [100][101], improves transverse aortic constriction (TAC)-induced mitochondrial damage and heart failure phenotypes, and suppresses mitochondrial proteomic changes [102]. Preclinical studies have shown that SS-31 by its anti-inflammatory and anti-oxidant effects alleviates sepsis-induced cardiac injury [103], improves failing heart function, ameliorates myocardial mitochondrial fragmentation [104], maintains myocardial mitochondrial integrity, and improves myocardial energy supply in rats with I/R [105][106]. In addition, SS-31 can also remove ROS, stabilize mitochondrial membrane potential, reduce cardiomyocyte apoptosis, and alleviate diastolic cardiomyopathy and myocardial toxicity induced by doxorubicin in rats [107][108]. In high-fat diet ApoE knockout mice, administration of SS-31 inhibits cholesterol uptake, which in turn inhibits foam cell formation and atherosclerosis progression [109]. In TAC-induced PAH mice model, SS-31 can inhibit the release of inflammatory factors and improve endothelial function by reducing the expression of pro-oxidant protein NOX1/NOX2, and reduce TAC-induced right ventricular systolic pressure [110]. These results indicate that SS-31 has great potential in the treatment of CVD.

5.2.4. MitoTEMPOL

MitoTEMPOL is a novel mitochondria-targeted antioxidant formed by the combination of the superoxide dismutase mimics Tempol and TPP+. MitoTEMPOL can scavenge mtROS, so it can alleviate the oxidative damage of mitochondria induced by mtROS under various pathological conditions, restore mitochondrial function, and maintain normal function of the body.
In CVD, MitoTEMPOL could improve ventricular arrhythmias, eliminate sudden cardiac death, and inhibit the expression of proteome remodeling and specific phosphoproteome alterations induced by chronic heart failure in guinea pigs, or preventing and reversing heart failure by removing ROS in the heart. Mechanistically, MitoTEMPOL repairs the disruption of normal coupling between cytoplasmic signaling and nuclear genetic programs, and restores the mitochondrial function, antioxidant enzymes, Ca2+ handling, and action potential repolarization disrupted by mtROS [111]. By clearing mtROS, MitoTEMPOL can inhibit the opening of the MPTP in the rat heart induced by nicotine, inhibit cardiac hypertrophy and cardiac fibrosis, and improve the sensitivity of rat myocardium to I/R injury caused by nicotine [112][113].

5.2.5. MitoSNO

Nitrosylation is one of the common post-translational modifications of proteins, which can regulate enzyme activity, subcellular localization and protein interactions. Studies have shown that nitrosylation modifications are also widely present in mitochondrial proteins and participate in the regulation of mitochondrial metabolic enzyme activities [114]. The reversible S-nitrosation modification is the main way that nitric oxide (NO) metabolism regulates mitochondrial physiology and pathology [114][115]. To further investigate the role of mitochondrial S-nitrosation, mitochondria-targeted MitoSNO has been formed by combining the NO donor SNAPNO (S-nitroso-N-acetylpenicillamine) and lipophilic TPP+. Driven by the positive charge of TPP+ and the mitochondrial inner membrane potential, MitoSNO accumulates rapidly in the mitochondria. Further study finds 13 S-nitrosation mitochondrial proteins, including aconitase, mitochondrial dehydrogenase and α-ketoglutarate dehydrogenase, whose activities are selectively and reversibly inhibited by S-nitrosation, indicating that S-nitrosation is reversibly involved in the regulation of enzyme activity of mitochondrial core metabolic enzymes [116].

5.3. Mitochondria-Targeted Gene Therapy Strategies

5.3.1. Mitochondria-Targeted Gene Editing Technologies

Because of the heterogeneity of mtDNA, mtDNA mutations need to accumulate to a threshold before they cause mitochondria-related diseases. For example, proband’s fibroblasts pathogenic variant m.8993T > G variant in MT-ATP6 subunit is 83%, and exhibits severe defects in mitochondrial cristae structure and decreased spare respiratory capacity in response to energy stress [117]. Therefore, it is a good idea to repair or excise mutated mtDNA by gene-editing technology to reduce the proportion of mutated mtDNA in the overall mtDNA. By combining an adeno-associated virus (AAV) 9-loaded TALEN (mitochondria-targeted TALEN, MitoTALEN) gene-editing tool with a mitochondrial guiding sequence, the m.5024C > T mutation site is successfully targeted to be cleaved, thus achieves the purpose of inhibiting the replication of the mutant gene. Six months after injection of this gene editing tool into the m.5024C > T mutant mouse model, the mtDNA mutation rate in mouse skeletal and heart muscle tissue decreases by 50%, most of the mutant genes are eliminated, and the mutation-induced reduction in transfer RNAAla levels is ameliorated [118].

5.3.2. Ectopic Expression of Mitochondrial Proteins

Ectopic expression of mitochondrial proteins is the process of constructing recombinant plasmids by combining mtDNA sequences encoding mitochondrial proteins and mitochondria-targeted sequences and integrating them into genomes through vectors (usually AAV) so that proteins originally expressed in mitochondria are sent to mitochondria through mitochondria-targeted peptides after being expressed in the nucleus and cytoplasm of cells to ameliorate metabolism dysfunction caused by impaired mitochondrial self-expression. Giovanni Manfredi et al. use ectopic expression of wild-type mtDNA encoding ATPase 6 of respiratory chain complex V in the nucleus and successfully import them into mitochondria, which significantly improve OXPHOS and ATP production in cells containing the ATP6 DNA 8993T > G mutation [119]. This technique allows the personalized design of recombinant plasmids to repair mitochondrial metabolic dysfunction caused by defects in mtDNA and is a potential mitochondria-targeted gene therapy strategy.

5.3.3. Mitochondrial Replacement Therapy (MRT) Technologies

The limitations of the conventional gene therapy approach make them fall short of expectations in the treatment of pathogenic mtDNA-defected diseases. MRT is the most effective technique for the prevention of hereditary mtDNA mutation disorders. This technique has been successfully tested in rhesus macaque monkeys [120] by transferring the nuclear genome of a mother carrying a pathogenic mtDNA mutation into the enucleated oocytes of another healthy mother by microinjection. However, the “three-parent” concept raised by this technique is controversial in terms of reproductive ethics. In addition, due to the possibility of selective replication and genetic drift, small amounts of residual maternally defective mtDNA may still gradually reach to pathogenic levels. Therefore, XiaoYan Fan et al. inject transmembrane peptide mRNA fused with autophagy receptor during nuclear transfer to reduce the carrying of defective mtDNA in reconstructed embryos by forcibly inducing the degradation of donor mitochondria [121]. To avoid ethical problems, scientists have developed co-culture technology to allow mtDNA-deficient cells to obtain autologous healthy mitochondria [122], or cytosolic fusion technology to fuse normal cells with mtDNA-deficient cells to form healthy hybrid cells to compensate for mitochondrial dysfunction in mtDNA-deficient cells [123].

This entry is adapted from the peer-reviewed paper 10.3390/pharmaceutics14122760

References

  1. Roth, G.A.; Mensah, G.A.; Johnson, C.O.; Addolorato, G.; Ammirati, E.; Baddour, L.M.; Barengo, N.C.; Beaton, A.Z.; Benjamin, E.J.; Benziger, C.P.; et al. Global burden of cardiovascular diseases and risk factors, 1990–2019: Update from the GBD 2019 study. J. Am. Coll. Cardiol. 2020, 76, 2982–3021.
  2. Jang, S.; Javadov, S. Association between ROS production, swelling and the respirasome integrity in cardiac mitochondria. Arch. Biochem. Biophys. 2017, 630, 1–8.
  3. Liu, J.; Wang, P.; Luo, J.; Huang, Y.; He, L.; Yang, H.; Li, Q.; Wu, S.; Zhelyabovska, O.; Yang, Q. Peroxisome proliferator-activated receptor β/δ activation in adult hearts facilitates mitochondrial function and cardiac performance under pressure-overload condition. Hypertension 2011, 57, 223–230.
  4. Caruso, P.; Dunmore, B.J.; Schlosser, K.; Schoors, S.; Dos Santos, C.; Perez-Iratxeta, C.; Lavoie, J.R.; Zhang, H.; Long, L.; Flockton, A.R.; et al. Identification of MicroRNA-124 as a major regulator of enhanced endothelial cell glycolysis in pulmonary arterial hypertension via PTBP1 (polypyrimidine tract binding protein) and pyruvate kinase M2. Circulation 2017, 136, 2451–2467.
  5. 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.
  6. Matsuura, Y.; Shimizu-Albergine, M.; Barnhart, S.; Kramer, F.; Hsu, C.-C.; Kothari, V.; Tang, J.; Gharib, S.A.; Kanter, J.E.; Abel, E.D.; et al. Diabetes suppresses glucose uptake and glycolysis in macrophages. Circ. Res. 2022, 130, 779–781.
  7. 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.
  8. Alberts, B.; Johnson, A.; Lewis, J.; Raff, M.; Roberts, K.; Walter, P. Energy conversion: Mitochondria and chloroplasts. In The Molecular Biology of the Cell, 4th ed.; Gibbs, S., Ed.; Garland Science: New York, NY, USA, 2002; p. 769.
  9. Sorrentino, V.; Menzies, K.J.; Auwerx, J. Repairing mitochondrial dysfunction in disease. Annu. Rev. Pharmacol. Toxicol. 2018, 58, 353–389.
  10. Fijalkowska, I.; Xu, W.; Comhair, S.A.A.; Janocha, A.J.; Mavrakis, L.A.; Krishnamachary, B.; Zhen, L.; Mao, T.; Richter, A.; Erzurum, S.C.; et al. Hypoxia inducible-factor1alpha regulates the metabolic shift of pulmonary hypertensive endothelial cells. Am. J. Pathol. 2010, 176, 1130–1138.
  11. Zhuang, W.; Lian, G.; Huang, B.; Du, A.; Gong, J.; Xiao, G.; Xu, C.; Wang, H.; Xie, L. CPT1 regulates the proliferation of pulmonary artery smooth muscle cells through the AMPK-p53-p21 pathway in pulmonary arterial hypertension. Mol. Cell. Biochem. 2019, 455, 169–183.
  12. Wall, V.Z.; Barnhart, S.; Kanter, J.E.; Kramer, F.; Shimizu-Albergine, M.; Adhikari, N.; Wight, T.N.; Hall, J.L.; Bornfeldt, K.E. Smooth muscle glucose metabolism promotes monocyte recruitment and atherosclerosis in a mouse model of metabolic syndrome. JCI Insight 2018, 3, e96544.
  13. Gupta, A.; Houston, B. A comprehensive review of the bioenergetics of fatty acid and glucose metabolism in the healthy and failing heart in nondiabetic condition. Heart Fail. Rev. 2017, 22, 825–842.
  14. Anderson, S.; Bankier, A.T.; Barrell, B.G.; de Bruijn, M.H.; Coulson, A.R.; Drouin, J.; Eperon, I.C.; Nierlich, D.P.; Roe, B.A.; Sanger, F.; et al. Sequence and organization of the human mitochondrial genome. Nature 1981, 290, 457–465.
  15. Andrews, R.M.; Kubacka, I.; Chinnery, P.F.; Lightowlers, R.N.; Turnbull, D.M.; Howell, N. Reanalysis and revision of the cambridge reference sequence for human mitochondrial DNA. Nat. Genet. 1999, 23, 147.
  16. Fontana, G.A.; Gahlon, H.L. Mechanisms of replication and repair in mitochondrial DNA deletion formation. Nucleic Acids Res. 2020, 48, 11244–11258.
  17. Kauppila, J.H.K.; Stewart, J.B. Mitochondrial DNA: Radically free of free-radical driven mutations. Biochim. Biophys. Acta 2015, 1847, 1354–1361.
  18. 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.
  19. Yu, H.; Zhang, F.; Yan, P.; Zhang, S.; Lou, Y.; Geng, Z.; Li, Z.; Zhang, Y.; Xu, Y.; Lu, Y.; et al. LARP7 protects against heart failure by enhancing mitochondrial biogenesis. Circulation 2021, 143, 2007–2022.
  20. Zhuang, L.; Jia, K.; Chen, C.; Li, Z.; Zhao, J.; Hu, J.; Zhang, H.; Fan, Q.; Huang, C.; Xie, H.; et al. DYRK1B-STAT3 drives cardiac hypertrophy and heart failure by impairing mitochondrial bioenergetics. Circulation 2022, 145, 829–846.
  21. Kalkhoran, S.B.; Kriston-Vizi, J.; Hernandez-Resendiz, S.; Crespo-Avilan, G.E.; Rosdah, A.A.; Lees, J.G.; Da Costa, J.R.S.; Ling, N.X.Y.; Holien, J.K.; Samangouei, P.; et al. Hydralazine protects the heart against acute ischaemia/reperfusion injury by inhibiting Drp1-mediated mitochondrial fission. Cardiovasc. Res. 2022, 118, 282–294.
  22. 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.
  23. Chen, K.-H.; Dasgupta, A.; Lin, J.; Potus, F.; Bonnet, S.; Iremonger, J.; Fu, J.; Mewburn, J.; Wu, D.; Dunham-Snary, K.; et al. Epigenetic dysregulation of the dynamin-related Protein 1 Binding Partners MiD49 and MiD51 Increases Mitotic Mitochondrial Fission and promotes pulmonary arterial hypertension: Mechanistic and therapeutic implications. Circulation 2018, 138, 287–304.
  24. Xiong, P.Y.; Tian, L.; Dunham-Snary, K.J.; Chen, K.-H.; Mewburn, J.D.; Neuber-Hess, M.; Martin, A.; Dasgupta, A.; Potus, F.; Archer, S.L. Biventricular increases in mitochondrial fission mediator (MiD51) and proglycolytic pyruvate kinase (PKM2) isoform in experimental group 2 pulmonary hypertension-novel mitochondrial abnormalities. Front. Cardiovasc. Med. 2018, 5, 195.
  25. Pendin, D.; Filadi, R.; Pizzo, P. The concerted action of mitochondrial dynamics and positioning: New characters in cancer onset and progression. Front. Oncol. 2017, 7, 102.
  26. Wai, T.; Langer, T. Mitochondrial dynamics and metabolic regulation. Trends Endocrinol. Metab. 2016, 27, 105–117.
  27. Li, T.; Han, J.; Jia, L.; Hu, X.; Chen, L.; Wang, Y. PKM2 coordinates glycolysis with mitochondrial fusion and oxidative phosphorylation. Protein Cell 2019, 10, 583–594.
  28. Skulachev, V.P. Mitochondrial filaments and clusters as intracellular power-transmitting cables. Trends Biochem. Sci. 2001, 26, 23–29.
  29. Tondera, D.; Grandemange, S.; Jourdain, A.; Karbowski, M.; Mattenberger, Y.; Herzig, S.; Da Cruz, S.; Clerc, P.; Raschke, I.; Merkwirth, C.; et al. SLP-2 is required for stress-induced mitochondrial hyperfusion. EMBO J. 2009, 28, 1589–1600.
  30. Liu, X.; Zhao, X.; Li, X.; Lv, S.; Ma, R.; Qi, Y.; Abulikemu, A.; Duan, H.; Guo, C.; Li, Y.; et al. PM2.5 triggered apoptosis in lung epithelial cells through the mitochondrial apoptotic way mediated by a ROS-DRP1-mitochondrial fission axis. J. Hazard. Mater. 2020, 397, 122608.
  31. Chan, C.-M.; Sekar, P.; Huang, D.-Y.; Hsu, S.-H.; Lin, W.-W. Different effects of metformin and A769662 on sodium iodate-induced cytotoxicity in retinal pigment epithelial cells: Distinct actions on mitochondrial fission and respiration. Antioxidants 2020, 9, 1057.
  32. Murphy, E.; Liu, J. Mitochondrial calcium and reactive oxygen species in cardiovascular disease. Cardiovasc. Res. 2022, 2022, cvac134.
  33. Shankar, T.S.; Ramadurai, D.K.A.; Steinhorst, K.; Sommakia, S.; Badolia, R.; Thodou Krokidi, A.; Calder, D.; Navankasattusas, S.; Sander, P.; Kwon, O.S.; et al. Cardiac-specific deletion of voltage dependent anion channel 2 leads to dilated cardiomyopathy by altering calcium homeostasis. Nat. Commun. 2021, 12, 4583.
  34. Hong, Z.; Chen, K.-H.; Dasgupta, A.; Potus, F.; Dunham-Snary, K.; Bonnet, S.; Tian, L.; Fu, J.; Breuils-Bonnet, S.; Provencher, S.; et al. MicroRNA-138 and MicroRNA-25 down-regulate mitochondrial calcium uniporter, causing the pulmonary arterial hypertension cancer phenotype. Am. J. Respir. Crit. Care Med. 2017, 195, 515–529.
  35. Honrath, B.; Matschke, L.; Meyer, T.; Magerhans, L.; Perocchi, F.; Ganjam, G.K.; Zischka, H.; Krasel, C.; Gerding, A.; Bakker, B.M.; et al. SK2 channels regulate mitochondrial respiration and mitochondrial Ca2+ uptake. Cell Death Differ. 2017, 24, 761–773.
  36. Assali, E.A.; Jones, A.E.; Veliova, M.; Acín-Pérez, R.; Taha, M.; Miller, N.; Shum, M.; Oliveira, M.F.; Las, G.; Liesa, M.; et al. NCLX prevents cell death during adrenergic activation of the brown adipose tissue. Nat. Commun. 2020, 11, 3347.
  37. Oropeza-Almazán, Y.; Vázquez-Garza, E.; Chapoy-Villanueva, H.; Torre-Amione, G.; García-Rivas, G. Small interfering RNA targeting mitochondrial calcium uniporter improves cardiomyocyte cell viability in hypoxia/reoxygenation injury by reducing calcium overload. Oxid. Med. Cell. Longev. 2017, 2017, 5750897.
  38. Chang, J.-C.; Lien, C.-F.; Lee, W.-S.; Chang, H.-R.; Hsu, Y.-C.; Luo, Y.-P.; Jeng, J.-R.; Hsieh, J.-C.; Yang, K.-T. Intermittent hypoxia prevents myocardial mitochondrial Ca2+ overload and cell death during ischemia/reperfusion: The role of reactive oxygen species. Cells 2019, 8, 564.
  39. Zanou, N.; Dridi, H.; Reiken, S.; Imamura de Lima, T.; Donnelly, C.; de Marchi, U.; Ferrini, M.; Vidal, J.; Sittenfeld, L.; Feige, J.N.; et al. Acute RyR1 Ca2+ leak enhances NADH-linked mitochondrial respiratory capacity. Nat. Commun. 2021, 12, 7219.
  40. Tarasov, A.I.; Griffiths, E.J.; Rutter, G.A. Regulation of ATP production by mitochondrial Ca(2+). Cell Calcium 2012, 52, 28–35.
  41. McCormack, J.G.; Denton, R.M. The effects of calcium ions and adenine nucleotides on the activity of pig heart 2-oxoglutarate dehydrogenase complex. Biochem. J. 1979, 180, 533–544.
  42. Denton, R.M.; McCormack, J.G.; Edgell, N.J. Role of calcium ions in the regulation of intramitochondrial metabolism. Effects of Na+, Mg2+ and ruthenium red on the Ca2+-stimulated oxidation of oxoglutarate and on pyruvate dehydrogenase activity in intact rat heart mitochondria. Biochem. J. 1980, 190, 107–117.
  43. Vatrinet, R.; Leone, G.; de Luise, M.; Girolimetti, G.; Vidone, M.; Gasparre, G.; Porcelli, A.M. The α-ketoglutarate dehydrogenase complex in cancer metabolic plasticity. Cancer Metab. 2017, 5, 3.
  44. Patel, M.S.; Nemeria, N.S.; Furey, W.; Jordan, F. The pyruvate dehydrogenase complexes: Structure-based function and regulation. J. Biol. Chem. 2014, 289, 16615–16623.
  45. Sheeran, F.L.; Angerosa, J.; Liaw, N.Y.; Cheung, M.M.; Pepe, S. Adaptations in protein expression and regulated activity of pyruvate dehydrogenase multienzyme complex in human systolic heart failure. Oxid. Med. Cell. Longev. 2019, 2019, 4532592.
  46. Liu, T.; O’Rourke, B. Regulation of mitochondrial Ca2+ and its effects on energetics and redox balance in normal and failing heart. J. Bioenerg. Biomembr. 2009, 41, 127–132.
  47. Denton, R.M.; Richards, D.A.; Chin, J.G. Calcium ions and the regulation of NAD+-linked isocitrate dehydrogenase from the mitochondria of rat heart and other tissues. Biochem. J. 1978, 176, 899–906.
  48. Lukács, G.L.; Kapus, A.; Fonyó, A. Parallel measurement of oxoglutarate dehydrogenase activity and matrix free Ca2+ in fura-2-loaded heart mitochondria. FEBS Lett. 1988, 229, 219–223.
  49. Balaban, R.S.; Nemoto, S.; Finkel, T. Mitochondria, oxidants, and aging. Cell 2005, 120, 483–495.
  50. Tormos, K.V.; Anso, E.; Hamanaka, R.B.; Eisenbart, J.; Joseph, J.; Kalyanaraman, B.; Chandel, N.S. Mitochondrial complex III ROS regulate adipocyte differentiation. Cell Metab. 2011, 14, 537–544.
  51. Loh, K.; Deng, H.; Fukushima, A.; Cai, X.; Boivin, B.; Galic, S.; Bruce, C.; Shields, B.J.; Skiba, B.; Ooms, L.M.; et al. Reactive oxygen species enhance insulin sensitivity. Cell Metab. 2009, 10, 260–272.
  52. Mailloux, R.J.; Fu, A.; Robson-Doucette, C.; Allister, E.M.; Wheeler, M.B.; Screaton, R.; Harper, M.-E. Glutathionylation state of uncoupling protein-2 and the control of glucose-stimulated insulin secretion. J. Biol. Chem. 2012, 287, 39673–39685.
  53. Mailloux, R.J.; Bériault, R.; Lemire, J.; Singh, R.; Chénier, D.R.; Hamel, R.D.; Appanna, V.D. The tricarboxylic acid cycle, an ancient metabolic network with a novel twist. PLoS ONE 2007, 2, e690.
  54. Handy, D.E.; Loscalzo, J. Redox regulation of mitochondrial function. Antioxid. Redox Signal. 2012, 16, 1323–1367.
  55. Ristow, M.; Schmeisser, K. Mitohormesis: Promoting health and lifespan by increased levels of reactive oxygen species (ROS). Dose Response 2014, 12, 288–341.
  56. Kehrer, J.P.; Klotz, L.-O. Free radicals and related reactive species as mediators of tissue injury and disease: Implications for Health. Crit. Rev. Toxicol. 2015, 45, 765–798.
  57. Ochoa, C.D.; Wu, R.F.; Terada, L.S. ROS signaling and ER stress in cardiovascular disease. Mol. Aspects Med. 2018, 63, 18–29.
  58. Zhang, J.; Wang, X.; Vikash, V.; Ye, Q.; Wu, D.; Liu, Y.; Dong, W. ROS and ROS-mediated cellular signaling. Oxid. Med. Cell. Longev. 2016, 2016, 4350965.
  59. Simonneau, G.; Montani, D.; Celermajer, D.S.; Denton, C.P.; Gatzoulis, M.A.; Krowka, M.; Williams, P.G.; Souza, R. Haemodynamic definitions and updated clinical classification of pulmonary hypertension. Eur. Respir. J. 2019, 53, 1801913.
  60. D’Alessandro, A.; El Kasmi, K.C.; Plecitá-Hlavatá, L.; Ježek, P.; Li, M.; Zhang, H.; Gupte, S.A.; Stenmark, K.R. Hallmarks of pulmonary hypertension: Mesenchymal and inflammatory cell metabolic reprogramming. Antioxid. Redox Signal. 2018, 28, 230–250.
  61. Xu, W.; Janocha, A.J.; Erzurum, S.C. Metabolism in pulmonary hypertension. Annu. Rev. Physiol. 2021, 83, 551–576.
  62. Xu, W.; Comhair, S.A.A.; Chen, R.; Hu, B.; Hou, Y.; Zhou, Y.; Mavrakis, L.A.; Janocha, A.J.; Li, L.; Zhang, D.; et al. Integrative proteomics and phosphoproteomics in pulmonary arterial hypertension. Sci. Rep. 2019, 9, 18623.
  63. Zhao, Y.; Peng, J.; Lu, C.; Hsin, M.; Mura, M.; Wu, L.; Chu, L.; Zamel, R.; Machuca, T.; Waddell, T.; et al. Metabolomic heterogeneity of pulmonary arterial hypertension. PLoS ONE 2014, 9, e88727.
  64. Lee, M.H.; Sanders, L.; Kumar, R.; Hernandez-Saavedra, D.; Yun, X.; Ford, J.A.; Perez, M.J.; Mickael, C.; Gandjeva, A.; Koyanagi, D.E.; et al. Contribution of fatty acid oxidation to the pathogenesis of pulmonary hypertension. Am. J. Physiol. Lung Cell. Mol. Physiol. 2022, 323, L355–L371.
  65. Laine, M.T.; Laukontaus, S.J.; Sund, R.; Aho, P.S.; Kantonen, I.; Albäck, A.; Venermo, M. A population-based study of abdominal aortic aneurysm treatment in finland 2000 to 2014. Circulation 2017, 136, 1726–1734.
  66. Ouyang, M.; Wang, M.; Yu, B. Aberrant mitochondrial dynamics: An emerging pathogenic driver of abdominal aortic aneurysm. Cardiovasc. Ther. 2021, 2021, 6615400.
  67. Glass, C.K.; Witztum, J.L. Atherosclerosis. the road ahead. Cell 2001, 104, 503–516.
  68. Pircher, A.; Treps, L.; Bodrug, N.; Carmeliet, P. Endothelial cell metabolism: A novel player in atherosclerosis? Basic principles and therapeutic opportunities. Atherosclerosis 2016, 253, 247–257.
  69. Orihuela, R.; McPherson, C.A.; Harry, G.J. Microglial M1/M2 polarization and metabolic states. Br. J. Pharmacol. 2016, 173, 649–665.
  70. Xu, R.-H.; Liu, B.; Wu, J.-D.; Yan, Y.-Y.; Wang, J.-N. miR-143 is involved in endothelial cell dysfunction through suppression of glycolysis and correlated with atherosclerotic plaques formation. Eur. Rev. Med. Pharmacol. Sci. 2016, 20, 4063–4071.
  71. Liu, L.; Lu, Y.; Martinez, J.; Bi, Y.; Lian, G.; Wang, T.; Milasta, S.; Wang, J.; Yang, M.; Liu, G.; et al. Proinflammatory signal suppresses proliferation and shifts macrophage metabolism from Myc-dependent to HIF1α-dependent. Proc. Natl. Acad. Sci. USA 2016, 113, 1564–1569.
  72. Tawakol, A.; Singh, P.; Mojena, M.; Pimentel-Santillana, M.; Emami, H.; MacNabb, M.; Rudd, J.H.F.; Narula, J.; Enriquez, J.A.; Través, P.G.; et al. HIF-1α and PFKFB3 mediate a tight relationship between proinflammatory activation and anerobic metabolism in atherosclerotic macrophages. Arterioscler. Thromb. Vasc. Biol. 2015, 35, 1463–1471.
  73. Kumar, V.; Santhosh Kumar, T.R.; Kartha, C.C. Mitochondrial membrane transporters and metabolic switch in heart failure. Heart Fail. Rev. 2019, 24, 255–267.
  74. Bertero, E.; Maack, C. Metabolic remodelling in heart failure. Nat. Rev. Cardiol. 2018, 15, 457–470.
  75. Lopaschuk, G.D.; Karwi, Q.G.; Tian, R.; Wende, A.R.; Abel, E.D. Cardiac energy metabolism in heart failure. Circ. Res. 2021, 128, 1487–1513.
  76. Jiang, L.; Zhou, S.; Zhang, X.; Li, C.; Ji, S.; Mao, H.; Jiang, X. Mitochondrion-specific dendritic lipopeptide liposomes for targeted sub-cellular delivery. Nat. Commun. 2021, 12, 2390.
  77. Zhang, C.-X.; Cheng, Y.; Liu, D.-Z.; Liu, M.; Cui, H.; Zhang, B.; Mei, Q.-B.; Zhou, S.-Y. Mitochondria-targeted cyclosporin A delivery system to treat myocardial ischemia reperfusion injury of rats. J. Nanobiotechn. 2019, 17, 18.
  78. Liu, C.-S.; Chang, J.-C.; Kuo, S.-J.; Liu, K.-H.; Lin, T.-T.; Cheng, W.-L.; Chuang, S.-F. Delivering healthy mitochondria for the therapy of mitochondrial diseases and beyond. Int. J. Biochem. Cell Biol. 2014, 53, 141–146.
  79. Nakazato, I.; Okuno, M.; Zhou, C.; Itoh, T.; Tsutsumi, N.; Takenaka, M.; Arimura, S.-I. Targeted base editing in the mitochondrial genome of Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 2022, 119, e2121177119.
  80. Shi, H.; Tang, H.; Ai, W.; Zeng, Q.; Yang, H.; Zhu, F.; Wei, Y.; Feng, R.; Wen, L.; Pu, P.; et al. Schisandrin B antagonizes cardiotoxicity induced by pirarubicin by inhibiting mitochondrial permeability transition pore (mPTP) opening and decreasing cardiomyocyte apoptosis. Front. Pharmacol. 2021, 12, 733805.
  81. Huang, P.; Wu, S.-P.; Wang, N.; Seto, S.; Chang, D. Hydroxysafflor yellow A alleviates cerebral ischemia reperfusion injury by suppressing apoptosis via mitochondrial permeability transition pore. Phytomedicine 2021, 85, 153532.
  82. Godoy, J.A.; Rios, J.A.; Picón-Pagès, P.; Herrera-Fernández, V.; Swaby, B.; Crepin, G.; Vicente, R.; Fernández-Fernández, J.M.; Muñoz, F.J. Mitostasis, calcium and free radicals in health, aging and neurodegeneration. Biomolecules 2021, 11, 1012.
  83. Xu, H.-X.; Cui, S.-M.; Zhang, Y.-M.; Ren, J. Mitochondrial Ca2+ regulation in the etiology of heart failure: Physiological and pathophysiological implications. Acta Pharmacol. Sin. 2020, 41, 1301–1309.
  84. Parodi-Rullán, R.M.; Chapa-Dubocq, X.R.; Javadov, S. Acetylation of mitochondrial proteins in the heart: The role of SIRT3. Front. Physiol. 2018, 9, 1094.
  85. Yang, W.; Nagasawa, K.; Münch, C.; Xu, Y.; Satterstrom, K.; Jeong, S.; Hayes, S.D.; Jedrychowski, M.P.; Vyas, F.S.; Zaganjor, E.; et al. Mitochondrial sirtuin network reveals dynamic SIRT3-dependent deacetylation in response to membrane depolarization. Cell 2016, 167, 985–1000.e21.
  86. Jin, J.-Y.; Wei, X.-X.; Zhi, X.-L.; Wang, X.-H.; Meng, D. Drp1-dependent mitochondrial fission in cardiovascular disease. Acta Pharmacol. Sin. 2021, 42, 655–664.
  87. Yang, J.; Suo, H.; Song, J. Protective role of mitoquinone against impaired mitochondrial homeostasis in metabolic syndrome. Crit. Rev. Food Sci. Nutr. 2021, 61, 3857–3875.
  88. Pandi-Perumal, S.R.; Trakht, I.; Srinivasan, V.; Spence, D.W.; Maestroni, G.J.M.; Zisapel, N.; Cardinali, D.P. Physiological effects of melatonin: Role of melatonin receptors and signal transduction pathways. Prog. Neurobiol. 2008, 85, 335–353.
  89. Venegas, C.; García, J.A.; Escames, G.; Ortiz, F.; López, A.; Doerrier, C.; García-Corzo, L.; López, L.C.; Reiter, R.J.; Acuña-Castroviejo, D. Extrapineal melatonin: Analysis of its subcellular distribution and daily fluctuations. J. Pineal Res. 2012, 52, 217–227.
  90. Nduhirabandi, F.; Maarman, G.J. Melatonin in heart failure: A promising therapeutic strategy? Molecules 2018, 23, 1819.
  91. Zhou, H.; Ma, Q.; Zhu, P.; Ren, J.; Reiter, R.J.; Chen, Y. Protective role of melatonin in cardiac ischemia-reperfusion injury: From pathogenesis to targeted therapy. J. Pineal Res. 2018, 64, e12471.
  92. Xia, L.; Sun, C.; Zhu, H.; Zhai, M.; Zhang, L.; Jiang, L.; Hou, P.; Li, J.; Li, K.; Liu, Z.; et al. Melatonin protects against thoracic aortic aneurysm and dissection through SIRT1-dependent regulation of oxidative stress and vascular smooth muscle cell loss. J. Pineal Res. 2020, 69, e12661.
  93. Duan, L.; Li, S.; Wang, L.; Jing, Y.; Li, G.; Sun, Y.; Sun, W.; Li, Y.; Zhao, L.; Xin, S. Melatonin plays a critical protective role in nicotine-related abdominal aortic aneurysm. Front. Physiol. 2020, 11, 866.
  94. Ajoolabady, A.; Bi, Y.; McClements, D.J.; Lip, G.Y.H.; Des Richardson, R.; Reiter, R.J.; Klionsky, D.J.; Ren, J. Melatonin-based therapeutics for atherosclerotic lesions and beyond: Focusing on macrophage mitophagy. Pharmacol. Res. 2022, 176, 106072.
  95. Zhang, J.; Lu, X.; Liu, M.; Fan, H.; Zheng, H.; Zhang, S.; Rahman, N.; Wołczyński, S.; Kretowski, A.; Li, X. Melatonin inhibits inflammasome-associated activation of endothelium and macrophages attenuating pulmonary arterial hypertension. Cardiovasc. Res. 2020, 116, 2156–2169.
  96. MacLean, M.R. Melatonin: Shining some light on pulmonary hypertension. Cardiovasc. Res. 2020, 116, 2036–2037.
  97. Whitson, J.A.; Bitto, A.; Zhang, H.; Sweetwyne, M.T.; Coig, R.; Bhayana, S.; Shankland, E.G.; Wang, L.; Bammler, T.K.; Mills, K.F.; et al. SS-31 and NMN: Two paths to improve metabolism and function in aged hearts. Aging Cell 2020, 19, e13213.
  98. Chiao, Y.A.; Zhang, H.; Sweetwyne, M.; Whitson, J.; Ting, Y.S.; Basisty, N.; Pino, L.K.; Quarles, E.; Nguyen, N.-H.; Campbell, M.D.; et al. Late-life restoration of mitochondrial function reverses cardiac dysfunction in old mice. Elife 2020, 9, e55513.
  99. Zhang, H.; Alder, N.N.; Wang, W.; Szeto, H.; Marcinek, D.J.; Rabinovitch, P.S. Reduction of elevated proton leak rejuvenates mitochondria in the aged cardiomyocyte. Elife 2020, 9, e60827.
  100. Whitson, J.A.; Martín-Pérez, M.; Zhang, T.; Gaffrey, M.J.; Merrihew, G.E.; Huang, E.; White, C.C.; Kavanagh, T.J.; Qian, W.-J.; Campbell, M.D.; et al. Elamipretide (SS-31) treatment attenuates age-associated post-translational modifications of heart proteins. Geroscience 2021, 43, 2395–2412.
  101. Whitson, J.A.; Johnson, R.; Wang, L.; Bammler, T.K.; Imai, S.-I.; Zhang, H.; Fredrickson, J.; Latorre-Esteves, E.; Bitto, A.; MacCoss, M.J.; et al. Age-related disruption of the proteome and acetylome in mouse hearts is associated with loss of function and attenuated by elamipretide (SS-31) and nicotinamide mononucleotide (NMN) treatment. Geroscience 2022, 44, 1621–1639.
  102. Dai, D.-F.; Hsieh, E.J.; Chen, T.; Menendez, L.G.; Basisty, N.B.; Tsai, L.; Beyer, R.P.; Crispin, D.A.; Shulman, N.J.; Szeto, H.H.; et al. Global proteomics and pathway analysis of pressure-overload-induced heart failure and its attenuation by mitochondrial-targeted peptides. Circ. Heart Fail. 2013, 6, 1067–1076.
  103. Liu, Y.; Yang, W.; Sun, X.; Xie, L.; Yang, Y.; Sang, M.; Jiao, R. SS31 ameliorates sepsis-induced heart Injury by inhibiting oxidative stress and inflammation. Inflammation 2019, 42, 2170–2180.
  104. Aimo, A.; Castiglione, V.; Borrelli, C.; Saccaro, L.F.; Franzini, M.; Masi, S.; Emdin, M.; Giannoni, A. Oxidative stress and inflammation in the evolution of heart failure: From pathophysiology to therapeutic strategies. Eur. J. Prev. Cardiol. 2020, 27, 494–510.
  105. Machiraju, P.; Wang, X.; Sabouny, R.; Huang, J.; Zhao, T.; Iqbal, F.; King, M.; Prasher, D.; Lodha, A.; Jimenez-Tellez, N.; et al. SS-31 peptide reverses the mitochondrial fragmentation present in fibroblasts from patients with DCMA, a mitochondrial cardiomyopathy. Front. Cardiovasc. Med. 2019, 6, 167.
  106. 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.
  107. Zhang, L.; Feng, M.; Wang, X.; Zhang, H.; Ding, J.; Cheng, Z.; Qian, L. Peptide Szeto-Schiller 31 ameliorates doxorubicin-induced cardiotoxicity by inhibiting the activation of the p38 MAPK signaling pathway. Int. J. Mol. Med. 2021, 47, 63.
  108. Yeh, J.-N.; Sung, P.-H.; Chiang, J.Y.; Sheu, J.-J.; Huang, C.-R.; Chu, Y.-C.; Chua, S.; Yip, H.-K. Early treatment with combination of SS31 and entresto effectively preserved the heart function in doxorubicin-induced dilated cardiomyopathic rat. Biomed. Pharmacother. 2021, 141, 111886.
  109. Zhang, M.; Zhao, H.; Cai, J.; Li, H.; Wu, Q.; Qiao, T.; Li, K. Chronic administration of mitochondrion-targeted peptide SS-31 prevents atherosclerotic development in ApoE knockout mice fed Western diet. PLoS ONE 2017, 12, e0185688.
  110. Lu, H.-I.; Huang, T.-H.; Sung, P.-H.; Chen, Y.-L.; Chua, S.; Chai, H.-Y.; Chung, S.-Y.; Liu, C.-F.; Sun, C.-K.; Chang, H.-W.; et al. Administration of antioxidant peptide SS-31 attenuates transverse aortic constriction-induced pulmonary arterial hypertension in mice. Acta Pharmacol. Sin. 2016, 37, 589–603.
  111. Dey, S.; DeMazumder, D.; Sidor, A.; Foster, D.B.; O’Rourke, B. Mitochondrial ROS drive sudden cardiac death and chronic proteome remodeling in heart failure. Circ. Res. 2018, 123, 356–371.
  112. Fang, X.; Wang, H.; Han, D.; Xie, E.; Yang, X.; Wei, J.; Gu, S.; Gao, F.; Zhu, N.; Yin, X.; et al. Ferroptosis as a target for protection against cardiomyopathy. Proc. Natl. Acad. Sci. USA 2019, 116, 2672–2680.
  113. Ramalingam, A.; Budin, S.B.; Mohd Fauzi, N.; Ritchie, R.H.; Zainalabidin, S. Targeting mitochondrial reactive oxygen species-mediated oxidative stress attenuates nicotine-induced cardiac remodeling and dysfunction. Sci. Rep. 2021, 11, 13845.
  114. Moncada, S.; Erusalimsky, J.D. Does nitric oxide modulate mitochondrial energy generation and apoptosis? Nat. Rev. Mol. Cell Biol. 2002, 3, 214–220.
  115. Dahm, C.C.; Moore, K.; Murphy, M.P. Persistent S-nitrosation of complex I and other mitochondrial membrane proteins by S-nitrosothiols but not nitric oxide or peroxynitrite: Implications for the interaction of nitric oxide with mitochondria. J. Biol. Chem. 2006, 281, 10056–10065.
  116. Chouchani, E.T.; Hurd, T.R.; Nadtochiy, S.M.; Brookes, P.S.; Fearnley, I.M.; Lilley, K.S.; Smith, R.A.J.; Murphy, M.P. Identification of S-nitrosated mitochondrial proteins by S-nitrosothiol difference in gel electrophoresis (SNO-DIGE): Implications for the regulation of mitochondrial function by reversible S-nitrosation. Biochem. J. 2010, 430, 49–59.
  117. Uittenbogaard, M.; Brantner, C.A.; Fang, Z.; Wong, L.-J.C.; Gropman, A.; Chiaramello, A. Novel insights into the functional metabolic impact of an apparent de novo m.8993TG variant in the MT-ATP6 gene associated with maternally inherited form of Leigh Syndrome. Mol. Genet. Metab. 2018, 124, 71–81.
  118. Bacman, S.R.; Kauppila, J.H.K.; Pereira, C.V.; Nissanka, N.; Miranda, M.; Pinto, M.; Williams, S.L.; Larsson, N.-G.; Stewart, J.B.; Moraes, C.T. MitoTALEN reduces mutant mtDNA load and restores tRNAAla levels in a mouse model of heteroplasmic mtDNA mutation. Nat. Med. 2018, 24, 1696–1700.
  119. Manfredi, G.; Fu, J.; Ojaimi, J.; Sadlock, J.E.; Kwong, J.Q.; Guy, J.; Schon, E.A. Rescue of a deficiency in ATP synthesis by transfer of MTATP6, a mitochondrial DNA-encoded gene, to the nucleus. Nat. Genet. 2002, 30, 394–399.
  120. Tachibana, M.; Sparman, M.; Sritanaudomchai, H.; Ma, H.; Clepper, L.; Woodward, J.; Li, Y.; Ramsey, C.; Kolotushkina, O.; Mitalipov, S. Mitochondrial gene replacement in primate offspring and embryonic stem cells. Nature 2009, 461, 367–372.
  121. Fan, X.-Y.; Guo, L.; Chen, L.-N.; Yin, S.; Wen, J.; Li, S.; Ma, J.-Y.; Jing, T.; Jiang, M.-X.; Sun, X.-H.; et al. Reduction of mtDNA heteroplasmy in mitochondrial replacement therapy by inducing forced mitophagy. Nat. Biomed. Eng. 2022, 6, 339–350.
  122. Kitani, T.; Kami, D.; Kawasaki, T.; Nakata, M.; Matoba, S.; Gojo, S. Direct human mitochondrial transfer: A novel concept based on the endosymbiotic theory. Transplant. Proc. 2014, 46, 1233–1236.
  123. Wilkins, H.M.; Carl, S.M.; Swerdlow, R.H. Cytoplasmic hybrid (cybrid) cell lines as a practical model for mitochondriopathies. Redox Biol. 2014, 2, 619–631.
More
This entry is offline, you can click here to edit this entry!
Video Production Service