Next Article in Journal
From Gasotransmitter to Immunomodulator: The Emerging Role of Hydrogen Sulfide in Macrophage Biology
Next Article in Special Issue
Intraoperative Hemi-Diaphragm Electrical Stimulation Demonstrates Attenuated Mitochondrial Function without Change in Oxidative Stress in Cardiothoracic Surgery Patients
Previous Article in Journal
Influence of Redox and Dopamine Regulation in Cocaine-Induced Phenotypes Using Drosophila
Previous Article in Special Issue
Reverse Electron Transport at Mitochondrial Complex I in Ischemic Stroke, Aging, and Age-Related Diseases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Targeting Mitochondrial Oxidative Stress as a Strategy to Treat Aging and Age-Related Diseases

1
Division of Life Sciences, College of Life Sciences and Bioengineering, Incheon National University, Incheon 22012, Republic of Korea
2
Convergence Research Center for Insect Vectors, Incheon National University, Incheon 22012, Republic of Korea
3
The Future Life & Society Research Center, Chonnam National University, Gwangju 61186, Republic of Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Antioxidants 2023, 12(4), 934; https://doi.org/10.3390/antiox12040934
Submission received: 6 March 2023 / Revised: 11 April 2023 / Accepted: 13 April 2023 / Published: 15 April 2023
(This article belongs to the Special Issue Mitochondrial Oxidative Stress in Aging and Disease)

Abstract

:
Mitochondria are one of the organelles undergoing rapid alteration during the senescence process. Senescent cells show an increase in mitochondrial size, which is attributed to the accumulation of defective mitochondria, which causes mitochondrial oxidative stress. Defective mitochondria are also targets of mitochondrial oxidative stress, and the vicious cycle between defective mitochondria and mitochondrial oxidative stress contributes to the onset and development of aging and age-related diseases. Based on the findings, strategies to reduce mitochondrial oxidative stress have been suggested for the effective treatment of aging and age-related diseases. In this article, we discuss mitochondrial alterations and the consequent increase in mitochondrial oxidative stress. Then, the causal role of mitochondrial oxidative stress on aging is investigated by examining how aging and age-related diseases are exacerbated by induced stress. Furthermore, we assess the importance of targeting mitochondrial oxidative stress for the regulation of aging and suggest different therapeutic strategies to reduce mitochondrial oxidative stress. Therefore, this review will not only shed light on a new perspective on the role of mitochondrial oxidative stress in aging but also provide effective therapeutic strategies for the treatment of aging and age-related diseases through the regulation of mitochondrial oxidative stress.

1. Introduction

When somatic cells reach a certain number of mitotic levels, they begin to lose their ability to proliferate, which is one of the hallmarks of senescence [1]. Along with losing the capacity to divide, senescent cells show dramatic changes, such as enlarged and flattened cell morphology, increased production of reactive oxygen species (ROS), the accumulation of consequent ROS-mediated damage derivatives (e.g., lipofuscin and granules), and the senescence-associated secretory phenotype (SASP) [2]. Senescent cells accumulate with age, negatively impacting regenerative capacity and creating a proinflammatory environment conducive to the onset and development of aging and age-related diseases [2]. This causal relationship is supported by the finding that the selective removal of senescent cells in vivo reduces inflammation and improves immune system function, slowing the development of aging and thereby extending lifespan [3]. Furthermore, senolytic therapies that eliminate senescent cells prevent age-related bone loss and fragility, supporting the causal link between senescence and aging [4].
Changes in organelle morphology or function are another characteristic of senescence, among which mitochondrial degeneration is most prominent [5,6]. Mitochondria exhibit structural changes such as significant increases in volume and size due to the buildup of defective mitochondria [7]. Defective mitochondria generate ROS as a byproduct of electron leakage from the electron transport complex (ETC) [8]. Not only are defective mitochondria ROS generators, but they are also targets of mitochondrial oxidative stress, which then boosts mitochondrial ROS production. Mitochondrial ROS generated by defective mitochondria deteriorate the morphology and function of organelles, consequently leading to aging and age-related diseases [9]. Therefore, strategies to reduce mitochondrial oxidative stress may be beneficial as therapeutic approaches to aging and age-related diseases [10,11]. The finding that treatment of senescent cells with ROS scavengers restored the senescent phenotype supports the usefulness of this strategy [11]. Mitochondrial oxidative stress is a major cause of senescence and the consequent development of age-related diseases, so a deeper comprehension of the mechanisms that target and control mitochondrial oxidative stress is needed.
This review discusses mitochondrial alterations and the consequent increase in mitochondrial oxidative stress and proposes ways to reduce mitochondrial oxidative stress to treat aging and age-related diseases. Using search terms including “mitochondrial alterations”, “mitochondrial oxidative stress”, and “mitochondrial ROS”, a thorough literature search was conducted in PubMed (a database of life sciences and medical journal articles). Based on previous and current studies derived from the literature, we provide a new perspective on the causal link between mitochondrial oxidative stress and aging and suggest potential therapeutic options for treating aging and age-related diseases.

2. Mitochondrial Alterations during the Process of Senescence and Aging

Mitochondria maintain their morphology, quality, and function through mitochondrial dynamics consisting of fusion and fission (Figure 1A). Mitochondrial fusion occurs in two steps. Mitofusin 1 (Mfn1) and mitofusin 2 (Mfn2), mitochondrial dynamin-like GTPases, fuse the outer mitochondrial membrane (OMM) [12,13] (Figure 1A). Then, OPA1 mitochondrial dynamin-like GTPase fuses the inner mitochondrial membrane (IMM). Mitochondrial fusion allows the mixing of the components of healthy and partially damaged mitochondria, resulting in a mitochondrial network with more uniform components [14]. Mitochondrial fission is regulated by receptors on OMM, including a mitochondrial dynamics protein of 49 kDa and 51 kDa (MiD49 and MiD51, respectively), mitochondrial fission factor (Mff), and fission 1 protein (FIS1) [15] (Figure 1A). The endoplasmic reticulum (ER) initiates mitochondrial fission by constricting the mitochondrial membrane, and then MiD49 and MiD51 recruit dynamin-related protein 1 (Drp1) to the mitochondrial surface [16,17] (Figure 1A). This binding allows higher-order Drp1 oligomers to form around the mitochondrial surface, leading to separation into two mitochondria [15,17]. A recent study subdivides mitochondrial fission into fission occurring in the mid-zone and periphery [18]. Mitochondrial fission at the mid-zone occurs when ER tubules contact mitochondria and constrict [18]. Then, Mff recruits Drp1 to scission sites. Mitochondrial fission at the periphery precedes lysosomal contact and is regulated by FIS1, which recruits Drp1 to scission sites [18]. Both types of mitochondrial fission are mediated by Drp1, inducing the formation of higher-order Drp1 oligomers around the scission site, splitting one mitochondrion into two mitochondria [18]. Mitochondrial fission at the mid-zone creates new mitochondria to provide necessary mitochondria during cell growth and division [18]. Mitochondrial fission at the periphery serves as quality control by isolating defective mitochondria from the mitochondrial network and allowing them to be eliminated by mitochondria-specific autophagy (mitophagy) [18]. Senescent cells show changes in mitochondrial morphology, such as increased mitochondrial mass and size [19,20]. Specifically, age-related lysosomal dysfunction prevents lysosomes from fusing with autophagosomes, limiting the efficient clearance of defective mitochondria via mitophagy and leading to the accumulation of defective mitochondria with aberrant and large morphology [21,22]. The causal role of senescence on mitochondrial morphology is corroborated by the finding that both senescent and H2O2-induced senescent cells express low levels of FIS1, resulting in an imbalance in mitochondrial fusion–fission [23] (Figure 1B and Table 1). This imbalance promotes the formation of large mitochondria with highly interconnected network structures [23] (Figure 1B and Table 1). This observation is supported by the discovery that senescent cells exhibit lower FIS1 and Drp1 expression, mediating the formation of large mitochondria and the resistance to oxidative stress through PTEN-induced putative kinase protein 1 (PINK1) [24]. Furthermore, FIS1 knockdown cells exhibit large mitochondria with concomitant senescence-related phenotypic changes [25]. Large mitochondria also limit the efficiency of clearing damaged mitochondria through mitophagy, reducing mitochondrial turnover. This phenomenon is evidenced by the finding that a 60% increase in mitochondrial size and a considerable rise in the percentage of large mitochondria were observed in cells from aged mice [26].
Mitochondrial homeostasis is primarily regulated by mitochondrial Ca2+ concentration. Ca2+ enters the mitochondria from the ER through mitochondrial porins known as voltage-dependent anion channels (VDACs) in the OMM (Figure 2A). Specifically, G protein-coupled receptor 75 (Gpr75) connects the inositol 1,4,5-triphosphate receptor (IP3R) in the ER with VDACs, facilitating Ca2+ influx from the ER into the mitochondria [27,28,29] (Figure 2A). Ca2+ passes via the mitochondrial calcium uniporter (MCU) in the IMM and is then taken up into the mitochondrial matrix. On the other hand, Ca2+ efflux occurs through channels that are distinct from those of Ca2+ influx. The Na+/Ca2+ exchanger (NCLX) and H+/Ca2+ exchanger (HCX) found in the IMM are responsible for Ca2+ efflux (Figure 2A). Senescence is characterized by Ca2+-overloaded mitochondria and is exacerbated by mitochondrial Ca2+-overload-induced mitochondrial oxidative stress [30,31,32,33] (Figure 2B and Table 1). Specifically, during oncogene-induced senescence, IP3R triggers a sustained increase in IP3R-mediated Ca2+ release [31]. Ca2+ then starts to be transported to the mitochondria through the VDAC/MCU channel [31] (Figure 2B and Table 1). Mitochondrial Ca2+ overload induces mitochondrial ROS generation and senescence [31]; the detailed mechanism will be further discussed in Section 3.
Table 1. A summary of senescence-associated mitochondrial alterations.
Table 1. A summary of senescence-associated mitochondrial alterations.
Mitochondrial AlterationOutcome(s)Experimental Model and References
Mitochondrial morphologyFormation of large mitochondria with highly interconnected network structuresMRC-5 human embryonic lung fibroblasts [19,20]
A considerable rise in the percentage of large mitochondriaC57/BL6 mice aged 30 months [26]
Mitochondrial Ca2+ homeostasisSenescent cells show a sustained increase in IP3R-mediated Ca2+ releaseHuman endometrium-derived stem cells [30]
Ca2+ then starts to be transported to the mitochondria through the VDAC/MCU channelHuman endometrial adenocarcinoma cells and WI38 human fibroblasts [31]
Mitochondria overloaded with Ca2+ exhibit increased electron leakage from the ETC and consequently generate mitochondrial ROSHuman endometrial adenocarcinoma cells and WI38 human fibroblasts [31]

3. Defective Mitochondria Are a Major Cause of Mitochondrial ROS Generation

More than 90% of oxygen is utilized by the mitochondria, and in complexes I and III of the ETC, 1–5% of oxygen is converted into short-lived superoxide anions (O2) [34]. Complexes I and III convert oxygen to superoxide anions (O2) in the mitochondrial matrix. In addition, complex III generates superoxide anions (O2) in the mitochondrial intermembrane space. The aging-related deterioration of complex I activity occurs due to decreased expression and increased oxidation of complex I [35,36]. Consistent with these findings, neurodegenerative diseases such as Parkinson’s disease (PD) have reduced complex I activity, leading to increased mitochondrial ROS production and complex I oxidation [37,38,39,40]. Complex I damaged by oxidative stress inefficiently transports electrons and subsequently increases electron leakage to oxygen, generating superoxide anions (O2) (Figure 3A and Table 2). Superoxide dismutase (SOD) converts superoxide anions (O2) to hydrogen peroxide (H2O2), a non-radical derivative that is relatively stable and permeable to the mitochondrial membrane. Then, hydrogen peroxide (H2O2) is partially reduced by the Fenton reaction and converted to a more harmful free radical, the hydroxyl radical (OH), which eventually causes severe mitochondrial oxidative stress [41] (Figure 3A). During the Fenton reaction, mitochondrial ROS is generated by mitochondrial iron that can be used for heme and iron–sulfur (Fe–S) cluster biosynthesis [42]. Mitochondrial dysfunction manifests as defects in heme and Fe–S cluster biosynthesis [43,44]. Alterations in iron homeostasis due to these defects lead to mitochondrial iron overload, resulting in the overproduction of free radicals via the Fenton reaction [45,46,47] (Table 2).
Mitochondrial Ca2+ overload is one of the main causes of mitochondrial ROS production. Specifically, mitochondrial Ca2+ overload stimulates the mitochondrial permeability transition (PT) and opens PT pores (mPTP), allowing ions and other solutes to move freely [48,49] (Figure 3B). The osmotic pressure of the mitochondrial matrix is increased by mPTP opening, which causes swelling of the mitochondria. Enlarged mitochondria readily lose cytochrome c that is loosely bound to the IMM [50,51,52,53] (Figure 3B and Table 2). Because cytochrome c plays a key role in electron transport in the ETC, the loss of cytochrome c impedes electron transport from complex I to IV, resulting in increased electron leakage (Figure 3B and Table 2). The leaked electrons react with oxygen to generate large amounts of mitochondrial ROS [50,51]. The induced increase in mitochondrial ROS generation causes mPTP to open more frequently and for longer periods of time [54,55]. The sustained opening of mPTP triggers a rapid decrease in the mitochondrial membrane potential and ATP synthesis. Furthermore, the continuous influx of ions and other solutes through mPTP causes electron leakage, which increases mitochondrial ROS production [53,55,56] (Figure 3B). Mitochondrial Ca2+ overload also increases hydrogen peroxide (H2O2) formation in mitochondria through the activation of α-glycerophosphate dehydrogenase (α-GPDH) located on the outer surface of the IMM [32]. This discovery is corroborated by the observation that the activation of α-GPDH by mitochondrial Ca2+ increases mitochondrial ROS production by providing more electrons to the ETC and creating favorable conditions for reverse electron transport [33].
Table 2. Defective mitochondria are a major cause of mitochondrial ROS generation.
Table 2. Defective mitochondria are a major cause of mitochondrial ROS generation.
Cause of Mitochondrial ROS GenerationOutcome(s)Experimental Model and ReferencesROS Related Information
Defective mitochondriaNeurodegenerative diseases such as PD have reduced complex I activity, leading to increased mitochondrial ROS production and complex I oxidationFibroblasts from the patient with PINK1 mutation [40]
  • ROS sources: mitochondrial ROS
  • Type of ROS: superoxide anions (O2), hydrogen peroxides (H2O2)
  • Enzymes involved in ROS generation: complex I
Mitochondrial dysfunction manifests as defects in heme and Fe–S cluster biosynthesis. Alterations in iron homeostasis due to these defects lead to mitochondrial iron overload, resulting in the overproduction of free radicals via the Fenton reactionChondrocyte C-20/A4 cell lines [47]
  • ROS sources: intracellular ROS
  • Type of ROS: total ROS
  • Enzymes involved in ROS generation: not specified
Mitochondrial Ca2+ overloadMitochondrial Ca2+ overload stimulates the mitochondrial permeability transition (PT) and opens PT pores (mPTP), allowing ions and other solutes to move freelyIsolated heart mitochondria from the bovine heart [48,49]
The osmotic pressure of the mitochondrial matrix is increased by mPTP opening, which causes swelling of the mitochondria. Enlarged mitochondria readily lose cytochrome c that is loosely bound to the IMMAstrocytes newborn C57BL/6 mice [53]
  • ROS sources: intracellular ROS
  • Type of ROS: total ROS
  • Enzymes involved in ROS generation: not specified
Loss of cytochrome c impedes electron transport from complex I to IV, resulting in increased electron leakage from the ETC. The leaked electrons react with oxygen to generate large amounts of mitochondrial ROSIsolated heart mitochondria from male Wistar rats [50]
  • ROS sources: mitochondrial ROS
  • Type of ROS: hydrogen peroxides (H2O2)
  • Enzymes involved in ROS generation: not specified

4. Vicious Feedback Loop between Mitochondrial Oxidative Stress and Senescence/Aging

ROS at physiologically low-to-moderate concentrations play biological roles in differentiation and proliferative responses [57,58]. For example, mild doses of oxidative stress induced by dietary restriction have been demonstrated to extend lifespan in various model organisms [59,60]. These results suggest that mild doses of oxidative stress can reduce chronic oxidative damage by increasing endogenous antioxidant defenses, triggering an adaptive response that enhances overall stress resistance [61]. Not only the level of ROS is important for cellular function, but the place where ROS is generated is also important. Mitochondrial ROS generation in complex I via reverse electron transport serves as a catalyst for preserving mitochondrial function and lengthening the lifespan in Drosophila [62]. However, in the same study, the mitochondrial ROS generation by preventing the activity of coenzyme Q (CoQ), which transports electrons from complex II to III, accelerates aging and shortens the lifespan in Drosophila [62]. These two conflicting results suggest that further studies are needed to understand how each ROS-producing site determines lifespan in Drosophila and whether each site plays a similar role in other organisms.
The physiological levels of ROS mediate important cellular functions, as shown in several studies, whereas the pathological levels of ROS cause irreversible damage to DNA, RNA, lipids, and proteins, which is a major contributor to aging and age-related diseases [63]. For example, accumulating oxidative damage over time activates the p53 and retinoblastoma protein pathways, resulting in persistent cell cycle arrest and senescence [9,64,65,66,67] (Figure 4). These findings are strengthened by recent studies showing that mitochondrial ROS activates a nicotinamide adenine dinucleotide (NAD+)-consuming enzyme called polyADP-ribose polymerase 1 [68,69]. Increased NAD+ consumption by polyADP-ribose polymerase 1 significantly reduces NAD+ levels [68]. As NAD+ is a crucial regulator allowing cells to respond to environmental changes, such as genotoxic factors and oxidative stress [70], a decrease in NAD+ levels leads to cellular dysfunction and aggravates age-related pathologies [71,72,73] (Figure 4). The role of mitochondrial oxidative stress on aging is further substantiated by the finding in the Caenorhabditis elegans (C. elegans) aging model caused by a mev-1 (complex II ortholog) mutation. The biochemical pathology of mev-1 mutants includes two-fold more mitochondrial ROS production than the wild type [74]. Increased mitochondrial ROS levels in mev-1 mutants led to premature age-dependent physiological alterations, including the accumulation of lipofuscin and protein carbonyl derivatives [75,76]. Moreover, the average and maximum lifespans of mev-1 mutants in oxygenated conditions were significantly decreased compared with those of the wild type [74]. Support for this phenomenon is seen in other studies targeting complex II. An iron chelator, deferoxamine, decreases complex II activity by inhibiting the conversion of Fe–S clusters in complex II [77]. Decreased complex II activity increases mitochondrial ROS production, resulting in senescence-like growth arrest [77,78]. Similar to these findings, transforming growth factor β-1 (TGF-β-1) activates mitochondrial ROS production by inactivating complex IV activity. The accumulation of mitochondrial oxidative stress through TGF-β-1-mediated inhibition directly causes degenerative changes [79,80,81]. The effect of mitochondrial oxidative stress on aging is further supported by findings in mice lacking superoxide dismutase 1 (SOD1), a protein found in both the mitochondrial matrix and the mitochondrial intermembrane space [82]. The absence of SOD1 increases superoxide anion (O2) production and consequent oxidative damage, resulting in symptoms commonly seen in aged mice, such as the premature loss of skeletal muscle mass [82].
Mitochondrial oxidative stress easily damages mitochondrial proteins involved in proteostasis, which is important for preserving and regulating protein quality in mitochondria [83]. Mitochondrial proteostasis consists of proteases that clearly damage mitochondrial proteins and chaperones that promote protein folding [84] (Figure 4). It has evolved to combat various mitochondrial stresses, including mitochondrial ROS-mediated damage. The impact of mitochondrial oxidative stress on mitochondrial proteostasis is shown in a PD mouse model [85]. The increase in ROS levels in the PD mouse model inactivates the Lon Peptidase 1, a mitochondrial protease, initiating an early event in PD pathogenesis [85]. The detrimental effects of mitochondrial oxidative stress on mitochondrial proteostasis are also shown in Alzheimer’s disease (AD) patients and transgenic AD mice [86]. Increased oxidative stress underlies the reduced activity of mitochondrial proteases, which aggravates AD pathogenesis by accumulating β-amyloid peptides in the mitochondria [86]. Furthermore, the prolonged disturbance of mitochondrial proteostasis by intracellular and mitochondrial ROS causes the accumulation of misfolded or aggregated mitochondrial proteins and triggers cellular senescence [87,88] (Figure 4).
Mitochondrial oxidative stress contributes to inflammation and proinflammatory secretory phenotypes (Figure 4). Specifically, superoxide anion (O2) reacts with SOD in the mitochondrial matrix to produce H2O2, which can pass through the OMM and react with cytoplasmic targets. This response activates the secretion of inflammatory cytokines (TNF-α and IL-1β) and proinflammatory SASP (IL-6 and IL-8) [89,90,91]. Mitochondrial ROS-mediated inflammatory responses exacerbate senescence, promoting chronic inflammation and age-related diseases in the long term [92]. Support for the causal link is evident in the discovery that mitochondria-targeted depletion inhibits the mitochondrial ROS production and secretion of key SASP factors such as IL-6 and IL-8 [93].
Mitochondrial DNA (mtDNA) is a circular chromosome found inside the mitochondria. mtDNA contains 37 genes encoding 2 rRNAs, 22 tRNAs, and 13 mitochondrial proteins, all essential for mitochondrial homeostasis [94]. Mitochondrial oxidative stress directly damages mtDNA, as mitochondrial ROS is produced in the mtDNA-containing mitochondrial matrix [95] (Figure 4). The increased production of mitochondrial ROS correlates linearly with the accumulation of mtDNA mutations [96]. Furthermore, mutations in mtDNA reduce the expression of essential proteins required for the ETC, leading to the amplification of mitochondrial oxidative stress [97]. This amplification results in a vicious cycle between mitochondrial oxidative stress and organelle degeneration [97]. The causal relationship between mtDNA mutation and aging is substantiated by the finding that the accumulation of mtDNA damage reduces ATP levels and triggers the recycling of β-amyloid toxicity, ultimately exacerbating neurodegeneration [98,99,100]. Support for this causal connection is evident in the finding that oxidative damage to mtDNA, but not to nuclear DNA, is inversely associated with maximal longevity in the heart and brain of mammals [101].

5. Targeting Mitochondrial Oxidative Stress as a Therapeutic Strategy for Aging and Age-Related Diseases

As described in Section 3 and Section 4, defective mitochondria are major sources of mitochondrial oxidative stress, and mitochondrial oxidative stress causes premature deterioration of tissue and organ function. The causal relationship suggests that the proper control of mitochondrial oxidative stress can be one of the effective therapeutic strategies for aging and age-related diseases. Here, we propose several therapeutic strategies targeting mitochondrial oxidative stress (Table 3).
Reducing mitochondrial oxidative stress with antioxidants represents an effective treatment strategy for aging. In mammals, an age-related decline in CoQ has been observed in the heart, liver, kidney, and skeletal muscle [102,103,104] (Figure 5A; green CoQ indicates CoQ deficiency). Cells deficient in CoQ exhibit electron leakage as electron transport in the ETC is hindered. Premature electron leakage from the ETC combines with oxygen to generate mitochondrial ROS [105] (Figure 5A). Considering that CoQ depletion triggers mitochondrial ROS production, mitochondria-targeted CoQ (MitoQ) treatment is tested in rat models of neurodegeneration [106]. As a mitochondria-specific antioxidant, MitoQ significantly lowers mitochondrial ROS production. Furthermore, MitoQ treatment is beneficial in treating neurodegeneration, showing reductions in mitochondrial swelling, cristae loss, and oxidative cell death [106] (Figure 5A; pink CoQ indicates high levels of CoQ). The benefit of CoQ supplementation in lowering mitochondrial oxidative stress is further supported by findings from an accelerated aging model using mice [107]. CoQ supplementation reduces mitochondrial ROS production by promoting the activity of complexes I and IV. Furthermore, a reduction in mitochondrial oxidative stress by CoQ supplementation delays the onset of age-related symptoms, indicating that therapeutics aimed at reducing mitochondrial oxidative stress might be effective [107]. In line with these findings, the treatment of senescent fibroblasts with N-acetylcysteine (NAC), a ROS scavenger, ameliorates senescence-associated phenotypes [11]. Moreover, antioxidant therapy with NAC ameliorates chondrocyte senescence and alleviates osteoarthritic phenotypes [108]. The justification of therapeutic strategies targeting mitochondrial oxidative stress is reinforced by experimental findings using SS-31, a cell-permeable antioxidant targeting the IMM [109]. Treatment with SS-31 boosts electron transport in the ETC, reducing the generation of hydroxyl and peroxynitrite free radicals in the mitochondria [109]. The reduction in mitochondrial oxidative stress by SS-31 has been implicated in delaying or reversing senescence in age-related diseases including AD and PD [110,111,112]. Additional experimental evidence for reducing mitochondrial oxidative stress with antioxidants is shown in findings using mice overexpressing mitochondria-targeted catalase (mCAT), an enzyme that breaks down hydrogen peroxide (H2O2) into H2O and O2 [113]. Two separate lines of mice overexpressing mCAT reduce hydrogen peroxide (H2O2) production and subsequent oxidative damage [114]. Moreover, these reductions delay cardiac pathology and cataract development, concomitant with a significant increase in the median and maximum lifespan [114].
The regulation of mitochondrial oxidative stress through the induction of mitophagy represents a promising and powerful strategy. Mitophagy, the selective degradation of damaged mitochondria through autophagy, is an essential mechanism to maintain mitochondrial homeostasis [115]. Specifically, the mitochondria damaged by high levels of mitochondrial ROS recruit PTEN-induced putative kinase protein 1 (PINK1) and ubiquitin ligase PARKIN, initiating the early phase of PINK1/PARKIN-dependent mitophagy [116] (Figure 5B). PINK1 activates PARKIN recruitment to the OMM for the ubiquitylation of OMM proteins in damaged mitochondria. The creation of polyubiquitin chains in OMM proteins recruits autophagy receptor proteins that interact with LC3 (microtubule-associated protein 1A/1B light chain 3) in the phagophore. Then, damaged mitochondria are eliminated by the subsequent formation of phagophores, autophagosomes, and autolysosomes, consequently resulting in reduced mitochondrial ROS production (Figure 5B). The significance of reducing mitochondrial oxidative stress by the activation of mitophagy has been validated by other findings. For example, the overexpression of PARKIN reduces mitochondrial oxidative stress with the protection of age-related loss of skeletal muscle in aged mice [117], whereas knocking out PARKIN accumulates markers of oxidative stress and deteriorates the contractile function of skeletal muscles [118]. Furthermore, PARKIN-mediated mitochondrial clearance suppresses mitochondrial ROS production and abrogates the development of senescence phenotypes such as pro-oxidant/inflammatory signaling and cyclin-dependent kinase inhibitors (p21 and p16) [93]. It is well known that PINK1 plays a key role in the interaction between OMM proteins and LC3 in the autophagosome, but recent studies have also shown that mitophagy can occur through a PINK1-independent mechanism. BNIP3 (BCL2 and adenovirus E1B 19-kDa-interacting protein 3) and NIP3-like protein X (NIX), which are localized to the OMM, serve as LC3 receptors, which contribute to the induction of mitophagy [119]. BNIP3 or NIX-mediated mitophagy also plays a crucial role in slowing and mitigating aging. For example, in Drosophila, the neuronal activation of BNIP3-mediated mitophagy delays systemic aging [120]. Moreover, NIX overexpression mitigates senescence by activating the degradation of mitochondrial proteins [121].
Consistent with these findings, mice that were given the mitophagy inducer trehalose were found to have reduced superoxide production and improved age-related atherosclerosis [122]. Similarly, therapy with the autophagy inducer lithium in C. elegans activates mitochondrial turnover with a decrease in mitochondrial oxidative stress [123]. A reduction in mitochondrial oxidative stress through lithium treatment consequently leads to an increase in the lifespan and healthspan without significantly changing death rates [123]. Furthermore, trehalose, a mitophagy inducer, has the effect of slowing down the aging process when it is administered to mice with age-related neurological symptoms induced by Atg7 knockdown [124]. The importance of mitophagy induction for regulating mitochondrial oxidative stress is further emphasized by findings related to nicotinamide riboside (NR) supplementation. NR supplementation raises NAD+ levels, which are compromised in aged animals [71,72,73]. Elevated NAD+ levels activate sirtuin 1, an NAD+-dependent histone/protein deacetylase that uses NAD+ to deacetylate forkhead box O3 and peroxisome proliferator-activated receptor co-activator 1-α (PGC-1α), transcriptional regulators that activate the genes involved in mitophagy [125,126]. Activated mitophagy then removes damaged mitochondria and reduces mitochondrial ROS levels, consequently alleviating disease pathology and prolonging lifespan in animal models [125,126]. Recent studies have provided new evidence that mitophagy induction through the modulation of ataxia telangiectasia mutagenesis (ATM) activity might be an effective treatment option for aging [127,128]. The inhibition of ATM activity increases V1–V0 assembly in the V-ATPase proton pump found at the lysosomal membrane, allowing re-acidification in lysosomes [127]. Acidified lysosomes activate mitophagy to clear damaged mitochondria and consequently lower mitochondrial ROS levels [127]. Reducing mitochondrial oxidative stress by simultaneously regulating ATM activity restores various senescence-associated phenotypes to the level of young cells [127,128].
Reducing mitochondrial oxidative stress by directly targeting genes that regulate the activity of ETC components could be an alternative therapeutic strategy. Rho-associated protein kinase (ROCK) controls mitochondrial ROS production by regulating the connections between Rac1b and cytochrome c [129] (Figure 5C). Specifically, ROCK activation phosphorylates Rac1b, making it easier for Rac1b to interact with cytochrome c. This interaction steals electrons from cytochrome c and causes a partial reduction in oxygen, triggering mitochondrial ROS production [129]. The inhibition of ROCK activity prevents Rac1b from intercepting electrons from cytochrome c, enabling efficient electron transport from complex III to IV [129] (Figure 5C). Then, efficient electron transport prompts complex IV activity, thereby reducing mitochondrial ROS production (Figure 5C). Reducing mitochondrial oxidative damage by regulating ROCK activity improves poor growth and the early beginning of senescent phenotypes [129,130]. The significance of regulating the activity of ETC components to reduce mitochondrial stress is also supported by other studies. Long-term caloric restriction (CR) upregulates the activity of complex IV, which partially offsets electron leak and reduces mitochondrial ROS production [131]. CR extends average and maximum lifespan by delaying aging and preventing the development of age-related symptoms [132,133]. Furthermore, epigallocatechin 3-gallate (EGCG), a CR mimic, also mitigates mitochondrial oxidative stress and restores the catalytic activity of complex I/ATP synthase [134]. EGCG treatment ameliorates the dysfunction of age-related immune disorders and extends the lifespan in various animal models [135,136,137,138].
Maintaining mitochondrial ROS levels through mitochondrial Ca2+ homeostasis also represents an effective therapeutic option. As described in Section 2, adequate levels of mitochondrial Ca2+ maintain mitochondrial homeostasis [139], whereas mitochondrial Ca2+ overload generates more mitochondrial ROS [140]. Since fine-tuning the mitochondrial Ca2+ concentrations can maintain mitochondrial ROS at an appropriate level [141], various strategies have been attempted to control mitochondrial Ca2+. For instance, myocardial reperfusion is an age-related disease manifested by reduced resistance to cardiac reperfusion damage [142]. The major cause of myocardial reperfusion is cardiac tissue damage due to the increased mitochondrial ROS production resulting from increased mitochondrial Ca2+ levels in cardiomyocytes [143]. Since Ca2+ transport from the cellular cytoplasm to the mitochondrial matrix is promoted by MCU in the IMM [144,145], ruthenium 360 (Ru360), a cell-permeable MCU inhibitor, was administered to the mitochondria in the reperfused heart. The inhibition of MCU by Ru360 reduces the percentage of mitochondria exhibiting Ca2+ overload and subsequently reduces the production of mitochondrial ROS [146] (Figure 5D). Subsequently, the pathological signs of mitochondrial swelling return to a steady state with reduced mitochondrial ROS production [146]. The significance of minimizing mitochondrial oxidative damage by maintaining appropriate levels of mitochondrial Ca2+ is further evidenced by the finding that the microRNA-mediated silencing of MCU shields cardiomyocytes from mitochondrial oxidative stress [147] (Figure 5D).
Reducing mitochondrial oxidative stress through sirtuin modulation can be an alternative therapy to control aging. Sirtuins respond to mitochondrial oxidative stress through deacetylating transcription factors that control antioxidant genes. In particular, sirtuin 1 deacetylates and activates PGC-1α, which increases the function of genes that can regulate oxidative stress, such as catalase, glutathione peroxidase, and manganese SOD [148]. The reduction in mitochondrial ROS levels by sirtuin 1 ameliorates the symptoms of age-related neurodegeneration by protecting cultured neurons from oxidative stress-mediated death [148]. These findings are strengthened by other studies showing that therapy with resveratrol, a sirtuin 1 activator, reduces H2O2-induced mitochondrial oxidative damage and protects against H2O2-induced cell death [149]. Furthermore, the overexpression of sirtuin 2 extends the shortened lifespan caused by hydrogen peroxide (H2O2) treatment, supporting the claim that reducing oxidative stress by regulating sirtuins is an effective strategy to control aging [150].
Table 3. Targeting mitochondrial oxidative stress as a therapeutic strategy for aging and age-related diseases.
Table 3. Targeting mitochondrial oxidative stress as a therapeutic strategy for aging and age-related diseases.
Therapeutic StrategyOutcome(s)Experimental Model and References
AntioxidantsMitochondria-targeted CoQ (MitoQ) treatment lowers mitochondrial ROS production, consequently resulting in reductions in mitochondrial swelling, cristae loss, and oxidative cell deathMale albino rats (Wistar strain) [106]
Treatment of senescent fibroblasts with N-acetylcysteine (NAC), a ROS scavenger, ameliorates senescence-associated phenotypesLaminopathy progeria fibroblasts [11]
Treatment with SS-31 boosts electron transport in the ETC, reducing the generation of free radicals in mitochondriaPrimary neurons from C57BL/6 mice [109]
MitophagyOverexpression of PARKIN reduces mitochondrial oxidative stress with the protection of age-related loss of skeletal muscle in aged miceSkeletal muscle of aged mice [117]
Knocking out PARKIN accumulates markers of oxidative stress and deteriorates the contractile function of skeletal musclesPARKIN knockout mice [118]
Mice given the mitophagy inducer trehalose showed reduced superoxide production and improved age-related atherosclerosisC57/BL6 mice aged 27–28 months [122]
Therapy with the autophagy inducer lithium in C. elegans activates mitochondrial turnover with a decrease in mitochondrial oxidative stressC. elegans [123]
Inhibition of ATM activity activates mitophagy and consequently lowers mitochondrial ROS levelsHuman fibroblasts [127]
Genes that regulate the activity of ETC componentsInhibition of the ROCK activity facilitates complex IV activity, thereby reducing mitochondrial ROS productionProgeria skin fibroblasts [129]
Long-term caloric restriction (CR) upregulates the activity of complex IV, which partially offsets electron leak and reduces mitochondrial ROS productionFemale Swiss Albino balb/c mice [131]
Mitochondrial Ca2+ homeostasisInhibition of MCU by Ru360 reduces the percentage of mitochondria exhibiting Ca2+ overload and subsequently reduces the production of mitochondrial ROSIsolated brain mitochondria from male Wistar rats [146]
MicroRNA-mediated silencing of MCU shields cardiomyocytes from mitochondrial oxidative stressRat cardiac myoblast H9c2 cell line [147]
SirtuinReduction in mitochondrial ROS levels by sirtuin 1 ameliorates symptoms of age-related neurodegeneration by protecting cultured neurons from oxidative stress-mediated deathPGC-1α null cell [148]
Overexpression of sirtuin 2 extends the shortened lifespan caused by hydrogen peroxide (H2O2) treatmentYeast cells [150]

6. Conclusions and Perspectives

Targeting mitochondrial oxidative stress as a therapeutic strategy to treat aging and age-related diseases has been investigated and discussed in several review articles. One review article discussed the impacts of oxidative stress on mitochondrial function and aging [151]. However, the discussion was limited to a specific cell type, liver cells. Another review article did not limit the discussion to a specific type of cell but extended it to various models of aging [152]. Although the origins and effects of mitochondrial oxidative stress have been thoroughly investigated in various aging models, specific methods to reduce mitochondrial oxidative stress have not been proposed. A recent review article highlighted mitochondrial oxidative stress as a major cause of aging and a crucial determinant of longevity [153]. In that article, the methods to reduce mitochondrial ROS were discussed, but the strategy was exclusively focused on mitochondrial antioxidants. Given the fact that mitochondrial oxidative stress is induced through multiple cellular signaling pathways, it is unclear whether therapeutic strategies focused only on antioxidants will be beneficial in treating aging and age-related disorders.
In this review, we investigated and discussed mitochondrial alterations and the consequent increase in mitochondrial oxidative stress. In addition, by examining the process through which mitochondrial oxidative stress progresses aging and aging-related diseases, we found that mitochondrial oxidative stress acts as a vicious feedback loop for aging. Here, we suggested mitochondrial oxidative stress as a potential target for aging and presented several therapeutic options aimed at reducing mitochondrial oxidative stress. Mitochondrial alterations increase with age and consequently induce mitochondrial oxidative stress, which has been effectively controlled using genetic and pharmacological approaches. Therefore, the optimal regulation of mitochondrial oxidative stress that does not depend on a single treatment strategy will effectively treat the onset and development of aging in which multiple signaling pathways are impaired.
As explored in this review, therapeutic approaches to reduce mitochondrial oxidative stress have proven to be an important factor in treating aging and age-related diseases. However, clinical trials using non-mitochondria-targeted antioxidants have shown that non-mitochondria-targeted antioxidant therapies are not effective in the treatment of aging and age-related diseases [154,155]. To complement these clinical findings, mitochondria-targeting antioxidants have recently been applied to various animal models, and there is growing evidence that mitochondria-targeting antioxidants have beneficial effects on aging and age-related diseases [156,157]. Therefore, research on how to effectively deliver genetic and pharmacological therapeutics targeting mitochondria will provide good therapeutic clues to break the vicious cycle leading to aging and age-related diseases.

Author Contributions

Writing—original draft preparation, Y.H.L., M.U.K., H.W.K., J.T.P. and S.C.P.; writing—review and editing, Y.H.L., M.U.K., J.T.P. and S.C.P.; data analysis, M.K.S., H.L., E.S.S. and S.K.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF), funded by the Ministry of Science, ICT, and Future Planning (NRF-2021R1A2C1004298). This research was also supported by the Priority Research Centers Program through the National Research Foundation of Korea (NRF), funded by the Ministry of Education (2020R1A6A1A0304195411).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the paper.

References

  1. Kumari, R.; Jat, P. Mechanisms of Cellular Senescence: Cell Cycle Arrest and Senescence Associated Secretory Phenotype. Front. Cell Dev. Biol. 2021, 9, 645593. [Google Scholar] [CrossRef] [PubMed]
  2. Hernandez-Segura, A.; Nehme, J.; Demaria, M. Hallmarks of Cellular Senescence. Trends Cell Biol. 2018, 28, 436–453. [Google Scholar] [CrossRef] [PubMed]
  3. Song, P.; An, J.; Zou, M.H. Immune Clearance of Senescent Cells to Combat Ageing and Chronic Diseases. Cells 2020, 9, 671. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Farr, J.N.; Xu, M.; Weivoda, M.M.; Monroe, D.G.; Fraser, D.G.; Onken, J.L.; Negley, B.A.; Sfeir, J.G.; Ogrodnik, M.B.; Hachfeld, C.M.; et al. Targeting cellular senescence prevents age-related bone loss in mice. Nat. Med. 2017, 23, 1072–1079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Bolden, J.E.; Lowe, S.W. 15—Cellular Senescence. In The Molecular Basis of Cancer, 4th ed.; Mendelsohn, J., Gray, J.W., Howley, P.M., Israel, M.A., Thompson, C.B., Eds.; Content Repository Only: Philadelphia, PA, USA, 2015; pp. 229–238.e222. [Google Scholar]
  6. Bouska, M.; Huang, K.; Kang, P.; Bai, H. Organelle aging: Lessons from model organisms. J. Genet. Genom. Yi Chuan Xue Bao 2019, 46, 171–185. [Google Scholar] [CrossRef]
  7. Boffoli, D.; Scacco, S.C.; Vergari, R.; Solarino, G.; Santacroce, G.; Papa, S. Decline with age of the respiratory chain activity in human skeletal muscle. Biochim. Biophys. Acta (BBA) Mol. Basis Dis. 1994, 1226, 73–82. [Google Scholar] [CrossRef]
  8. Hwang, E.; Yoon, G.; Kang, H. A comparative analysis of the cell biology of senescence and aging. Cell. Mol. Life Sci. 2009, 66, 2503–2524. [Google Scholar] [CrossRef]
  9. Giorgi, C.; Marchi, S.; Simoes, I.C.M.; Ren, Z.; Morciano, G.; Perrone, M.; Patalas-Krawczyk, P.; Borchard, S.; Jędrak, P.; Pierzynowska, K.; et al. Mitochondria and Reactive Oxygen Species in Aging and Age-Related Diseases. Int. Rev. Cell Mol. Biol. 2018, 340, 209–344. [Google Scholar] [CrossRef] [Green Version]
  10. Zhavoronkov, A.; Smit-McBride, Z.; Guinan, K.J.; Litovchenko, M.; Moskalev, A. Potential therapeutic approaches for modulating expression and accumulation of defective lamin A in laminopathies and age-related diseases. J. Mol. Med. 2012, 90, 1361–1389. [Google Scholar] [CrossRef] [Green Version]
  11. Richards, S.A.; Muter, J.; Ritchie, P.; Lattanzi, G.; Hutchison, C.J. The accumulation of un-repairable DNA damage in laminopathy progeria fibroblasts is caused by ROS generation and is prevented by treatment with N-acetyl cysteine. Hum. Mol. Genet. 2011, 20, 3997–4004. [Google Scholar] [CrossRef] [Green Version]
  12. Henze, K.; Martin, W. Evolutionary biology: Essence of mitochondria. Nature 2003, 426, 127–128. [Google Scholar] [CrossRef]
  13. Bernhardt, D.; Müller, M.; Reichert, A.S.; Osiewacz, H.D. Simultaneous impairment of mitochondrial fission and fusion reduces mitophagy and shortens replicative lifespan. Sci. Rep. 2015, 5, 7885. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Liu, Y.J.; McIntyre, R.L.; Janssens, G.E.; Houtkooper, R.H. Mitochondrial fission and fusion: A dynamic role in aging and potential target for age-related disease. Mech. Ageing Dev. 2020, 186, 111212. [Google Scholar] [CrossRef] [PubMed]
  15. Al Ojaimi, M.; Salah, A.; El-Hattab, A.W. Mitochondrial Fission and Fusion: Molecular Mechanisms, Biological Functions, and Related Disorders. Membranes 2022, 12, 893. [Google Scholar] [CrossRef] [PubMed]
  16. Palmer, C.S.; Osellame, L.D.; Laine, D.; Koutsopoulos, O.S.; Frazier, A.E.; Ryan, M.T. MiD49 and MiD51, new components of the mitochondrial fission machinery. EMBO Rep. 2011, 12, 565–573. [Google Scholar] [CrossRef]
  17. Palmer, C.S.; Elgass, K.D.; Parton, R.G.; Osellame, L.D.; Stojanovski, D.; Ryan, M.T. Adaptor proteins MiD49 and MiD51 can act independently of Mff and Fis1 in Drp1 recruitment and are specific for mitochondrial fission. J. Biol. Chem. 2013, 288, 27584–27593. [Google Scholar] [CrossRef] [Green Version]
  18. Kleele, T.; Rey, T.; Winter, J.; Zaganelli, S.; Mahecic, D.; Perreten Lambert, H.; Ruberto, F.P.; Nemir, M.; Wai, T.; Pedrazzini, T.; et al. Distinct fission signatures predict mitochondrial degradation or biogenesis. Nature 2021, 593, 435–439. [Google Scholar] [CrossRef]
  19. Lee, H.C.; Yin, P.H.; Chi, C.W.; Wei, Y.H. Increase in mitochondrial mass in human fibroblasts under oxidative stress and during replicative cell senescence. J. Biomed. Sci. 2002, 9, 517–526. [Google Scholar] [CrossRef]
  20. Passos, J.F.; Saretzki, G.; Ahmed, S.; Nelson, G.; Richter, T.; Peters, H.; Wappler, I.; Birket, M.J.; Harold, G.; Schaeuble, K.; et al. Mitochondrial dysfunction accounts for the stochastic heterogeneity in telomere-dependent senescence. PLoS Biol. 2007, 5, e110. [Google Scholar] [CrossRef] [Green Version]
  21. Guerrero-Navarro, L.; Jansen-Dürr, P.; Cavinato, M. Age-Related Lysosomal Dysfunctions. Cells 2022, 11, 1977. [Google Scholar] [CrossRef]
  22. Tran, M.; Reddy, P.H. Defective Autophagy and Mitophagy in Aging and Alzheimer’s Disease. Front. Neurosci. 2020, 14, 612757. [Google Scholar] [CrossRef] [PubMed]
  23. Yoon, Y.S.; Yoon, D.S.; Lim, I.K.; Yoon, S.H.; Chung, H.Y.; Rojo, M.; Malka, F.; Jou, M.J.; Martinou, J.C.; Yoon, G. Formation of elongated giant mitochondria in DFO-induced cellular senescence: Involvement of enhanced fusion process through modulation of Fis1. J. Cell. Physiol. 2006, 209, 468–480. [Google Scholar] [CrossRef] [PubMed]
  24. Mai, S.; Klinkenberg, M.; Auburger, G.; Bereiter-Hahn, J.; Jendrach, M. Decreased expression of Drp1 and Fis1 mediates mitochondrial elongation in senescent cells and enhances resistance to oxidative stress through PINK1. J. Cell Sci. 2010, 123, 917–926. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Lee, S.; Jeong, S.Y.; Lim, W.C.; Kim, S.; Park, Y.Y.; Sun, X.; Youle, R.J.; Cho, H. Mitochondrial fission and fusion mediators, hFis1 and OPA1, modulate cellular senescence. J. Biol. Chem. 2007, 282, 22977–22983. [Google Scholar] [CrossRef] [Green Version]
  26. Wilson, P.D.; Franks, L.M. The effect of age on mitochondrial ultrastructure. Gerontologia 1975, 21, 81–94. [Google Scholar] [CrossRef]
  27. Rizzuto, R.; Marchi, S.; Bonora, M.; Aguiari, P.; Bononi, A.; De Stefani, D.; Giorgi, C.; Leo, S.; Rimessi, A.; Siviero, R.; et al. Ca2+ transfer from the ER to mitochondria: When, how and why. Biochim. Biophys. Acta 2009, 1787, 1342–1351. [Google Scholar] [CrossRef] [Green Version]
  28. Xu, H.; Guan, N.; Ren, Y.L.; Wei, Q.J.; Tao, Y.H.; Yang, G.S.; Liu, X.Y.; Bu, D.F.; Zhang, Y.; Zhu, S.N. IP(3)R-Grp75-VDAC1-MCU calcium regulation axis antagonists protect podocytes from apoptosis and decrease proteinuria in an Adriamycin nephropathy rat model. BMC Nephrol. 2018, 19, 140. [Google Scholar] [CrossRef] [Green Version]
  29. Szabadkai, G.; Bianchi, K.; Várnai, P.; De Stefani, D.; Wieckowski, M.R.; Cavagna, D.; Nagy, A.I.; Balla, T.; Rizzuto, R. Chaperone-mediated coupling of endoplasmic reticulum and mitochondrial Ca2+ channels. J. Cell Biol. 2006, 175, 901–911. [Google Scholar] [CrossRef] [Green Version]
  30. Borodkina, A.V.; Shatrova, A.N.; Deryabin, P.I.; Griukova, A.A.; Abushik, P.A.; Antonov, S.M.; Nikolsky, N.N.; Burova, E.B. Calcium alterations signal either to senescence or to autophagy induction in stem cells upon oxidative stress. Aging 2016, 8, 3400–3418. [Google Scholar] [CrossRef] [Green Version]
  31. Wiel, C.; Lallet-Daher, H.; Gitenay, D.; Gras, B.; Le Calvé, B.; Augert, A.; Ferrand, M.; Prevarskaya, N.; Simonnet, H.; Vindrieux, D.; et al. Endoplasmic reticulum calcium release through ITPR2 channels leads to mitochondrial calcium accumulation and senescence. Nat. Commun. 2014, 5, 3792. [Google Scholar] [CrossRef] [Green Version]
  32. Tretter, L.; Adam-Vizi, V. High Ca2+ load promotes hydrogen peroxide generation via activation of α-glycerophosphate dehydrogenase in brain mitochondria. Free. Radic. Biol. Med. 2012, 53, 2119–2130. [Google Scholar] [CrossRef] [PubMed]
  33. Tretter, L.; Takacs, K.; Kövér, K.; Adam-Vizi, V. Stimulation of H2O2 generation by calcium in brain mitochondria respiring on alpha-glycerophosphate. J. Neurosci. Res. 2007, 85, 3471–3479. [Google Scholar] [CrossRef] [PubMed]
  34. Turrens, J.F. Mitochondrial formation of reactive oxygen species. J. Physiol. 2003, 552, 335–344. [Google Scholar] [CrossRef] [PubMed]
  35. Choksi, K.B.; Nuss, J.E.; Deford, J.H.; Papaconstantinou, J. Age-related alterations in oxidatively damaged proteins of mouse skeletal muscle mitochondrial electron transport chain complexes. Free. Radic. Biol. Med. 2008, 45, 826–838. [Google Scholar] [CrossRef] [Green Version]
  36. Choksi, K.B.; Boylston, W.H.; Rabek, J.P.; Widger, W.R.; Papaconstantinou, J. Oxidatively damaged proteins of heart mitochondrial electron transport complexes. Biochim. Biophys. Acta 2004, 1688, 95–101. [Google Scholar] [CrossRef] [Green Version]
  37. Papa, S.; De Rasmo, D. Complex I deficiencies in neurological disorders. Trends Mol. Med. 2013, 19, 61–69. [Google Scholar] [CrossRef]
  38. Morais, V.A.; Verstreken, P.; Roethig, A.; Smet, J.; Snellinx, A.; Vanbrabant, M.; Haddad, D.; Frezza, C.; Mandemakers, W.; Vogt-Weisenhorn, D.; et al. Parkinson’s disease mutations in PINK1 result in decreased Complex I activity and deficient synaptic function. EMBO Mol. Med. 2009, 1, 99–111. [Google Scholar] [CrossRef]
  39. Keeney, P.M.; Xie, J.; Capaldi, R.A.; Bennett, J.P., Jr. Parkinson’s disease brain mitochondrial complex I has oxidatively damaged subunits and is functionally impaired and misassembled. J. Neurosci. Off. J. Soc. Neurosci. 2006, 26, 5256–5264. [Google Scholar] [CrossRef] [Green Version]
  40. Piccoli, C.; Sardanelli, A.; Scrima, R.; Ripoli, M.; Quarato, G.; D’Aprile, A.; Bellomo, F.; Scacco, S.; De Michele, G.; Filla, A.; et al. Mitochondrial respiratory dysfunction in familiar parkinsonism associated with PINK1 mutation. Neurochem. Res. 2008, 33, 2565–2574. [Google Scholar] [CrossRef]
  41. Prousek, J. Fenton chemistry in biology and medicine. Pure Appl. Chem. 2007, 79, 2325–2338. [Google Scholar] [CrossRef]
  42. Braymer, J.J.; Lill, R. Iron-sulfur cluster biogenesis and trafficking in mitochondria. J. Biol. Chem. 2017, 292, 12754–12763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Veatch, J.R.; McMurray, M.A.; Nelson, Z.W.; Gottschling, D.E. Mitochondrial dysfunction leads to nuclear genome instability via an iron-sulfur cluster defect. Cell 2009, 137, 1247–1258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Rouault, T.A.; Tong, W.-H. Iron–sulphur cluster biogenesis and mitochondrial iron homeostasis. Nat. Rev. Mol. Cell Biol. 2005, 6, 345–351. [Google Scholar] [CrossRef] [PubMed]
  45. Mancardi, D.; Mezzanotte, M.; Arrigo, E.; Barinotti, A.; Roetto, A. Iron Overload, Oxidative Stress, and Ferroptosis in the Failing Heart and Liver. Antioxidants 2021, 10, 1864. [Google Scholar] [CrossRef] [PubMed]
  46. Chen, W.J.; Kung, G.P.; Gnana-Prakasam, J.P. Role of Iron in Aging Related Diseases. Antioxidants 2022, 11, 865. [Google Scholar] [CrossRef]
  47. Karim, A.; Bajbouj, K.; Shafarin, J.; Qaisar, R.; Hall, A.C.; Hamad, M. Iron Overload Induces Oxidative Stress, Cell Cycle Arrest and Apoptosis in Chondrocytes. Front. Cell Dev. Biol. 2022, 10, 821014. [Google Scholar] [CrossRef]
  48. Hunter, D.R.; Haworth, R.A. The Ca2+-induced membrane transition in mitochondria. I. The protective mechanisms. Arch. Biochem. Biophys. 1979, 195, 453–459. [Google Scholar] [CrossRef]
  49. Hunter, D.R.; Haworth, R.A.; Southard, J.H. Relationship between configuration, function, and permeability in calcium-treated mitochondria. J. Biol. Chem. 1976, 251, 5069–5077. [Google Scholar] [CrossRef]
  50. Pasdois, P.; Parker, J.E.; Griffiths, E.J.; Halestrap, A.P. The role of oxidized cytochrome c in regulating mitochondrial reactive oxygen species production and its perturbation in ischaemia. Biochem. J. 2011, 436, 493–505. [Google Scholar] [CrossRef] [Green Version]
  51. Markevich, N.I.; Hoek, J.B. Computational modeling analysis of mitochondrial superoxide production under varying substrate conditions and upon inhibition of different segments of the electron transport chain. Biochim. Biophys. Acta (BBA) Bioenerg. 2015, 1847, 656–679. [Google Scholar] [CrossRef] [Green Version]
  52. Tam, Z.Y.; Cai, Y.H.; Gunawan, R. Elucidating cytochrome C release from mitochondria: Insights from an in silico three-dimensional model. Biophys. J. 2010, 99, 3155–3163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Wong, S.S.; Li, R.H.; Stadlin, A. Oxidative stress induced by MPTP and MPP(+): Selective vulnerability of cultured mouse astrocytes. Brain Res. 1999, 836, 237–244. [Google Scholar] [CrossRef] [PubMed]
  54. Goodell, S.; Cortopassi, G. Analysis of oxygen consumption and mitochondrial permeability with age in mice. Mech. Ageing Dev. 1998, 101, 245–256. [Google Scholar] [CrossRef] [PubMed]
  55. Zorov, D.B.; Filburn, C.R.; Klotz, L.O.; Zweier, J.L.; Sollott, S.J. Reactive oxygen species (ROS)-induced ROS release: A new phenomenon accompanying induction of the mitochondrial permeability transition in cardiac myocytes. J. Exp. Med. 2000, 192, 1001–1014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Zorov, D.B.; Juhaszova, M.; Sollott, S.J. Mitochondrial ROS-induced ROS release: An update and review. Biochim. Biophys. Acta 2006, 1757, 509–517. [Google Scholar] [CrossRef] [Green Version]
  57. Mittler, R. ROS Are Good. Trends Plant Sci. 2017, 22, 11–19. [Google Scholar] [CrossRef] [Green Version]
  58. Finkel, T. Oxygen radicals and signaling. Curr. Opin. Cell Biol. 1998, 10, 248–253. [Google Scholar] [CrossRef]
  59. Zid, B.M.; Rogers, A.N.; Katewa, S.D.; Vargas, M.A.; Kolipinski, M.C.; Lu, T.A.; Benzer, S.; Kapahi, P. 4E-BP extends lifespan upon dietary restriction by enhancing mitochondrial activity in Drosophila. Cell 2009, 139, 149–160. [Google Scholar] [CrossRef] [Green Version]
  60. Schulz, T.J.; Zarse, K.; Voigt, A.; Urban, N.; Birringer, M.; Ristow, M. Glucose restriction extends Caenorhabditis elegans life span by inducing mitochondrial respiration and increasing oxidative stress. Cell Metab. 2007, 6, 280–293. [Google Scholar] [CrossRef] [Green Version]
  61. Ristow, M.; Zarse, K. How increased oxidative stress promotes longevity and metabolic health: The concept of mitochondrial hormesis (mitohormesis). Exp. Gerontol. 2010, 45, 410–418. [Google Scholar] [CrossRef]
  62. Scialò, F.; Sriram, A.; Fernández-Ayala, D.; Gubina, N.; Lõhmus, M.; Nelson, G.; Logan, A.; Cooper, H.M.; Navas, P.; Enríquez, J.A.; et al. Mitochondrial ROS Produced via Reverse Electron Transport Extend Animal Lifespan. Cell Metab. 2016, 23, 725–734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Checa, J.; Aran, J.M. Reactive Oxygen Species: Drivers of Physiological and Pathological Processes. J. Inflamm. Res. 2020, 13, 1057–1073. [Google Scholar] [CrossRef] [PubMed]
  64. Macip, S.; Igarashi, M.; Fang, L.; Chen, A.; Pan, Z.Q.; Lee, S.W.; Aaronson, S.A. Inhibition of p21-mediated ROS accumulation can rescue p21-induced senescence. EMBO J. 2002, 21, 2180–2188. [Google Scholar] [CrossRef]
  65. Passos, J.F.; Nelson, G.; Wang, C.; Richter, T.; Simillion, C.; Proctor, C.J.; Miwa, S.; Olijslagers, S.; Hallinan, J.; Wipat, A.; et al. Feedback between p21 and reactive oxygen production is necessary for cell senescence. Mol. Syst. Biol. 2010, 6, 347. [Google Scholar] [CrossRef]
  66. Luo, Y.; Zou, P.; Zou, J.; Wang, J.; Zhou, D.; Liu, L. Autophagy regulates ROS-induced cellular senescence via p21 in a p38 MAPKα dependent manner. Exp. Gerontol. 2011, 46, 860–867. [Google Scholar] [CrossRef] [Green Version]
  67. Takahashi, A.; Ohtani, N.; Yamakoshi, K.; Iida, S.; Tahara, H.; Nakayama, K.; Nakayama, K.I.; Ide, T.; Saya, H.; Hara, E. Mitogenic signalling and the p16INK4a-Rb pathway cooperate to enforce irreversible cellular senescence. Nat. Cell Biol. 2006, 8, 1291–1297. [Google Scholar] [CrossRef]
  68. Zhang, F.; Xie, R.; Munoz, F.M.; Lau, S.S.; Monks, T.J. PARP-1 Hyperactivation and Reciprocal Elevations in Intracellular Ca2+ During ROS-Induced Nonapoptotic Cell Death. Toxicol. Sci. 2014, 140, 118–134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Liu, X.; Kang, J.; Wang, H.; Huang, T. Mitochondrial ROS contribute to oridonin-induced HepG2 apoptosis through PARP activation. Oncol. Lett. 2018, 15, 2881–2888. [Google Scholar] [CrossRef] [Green Version]
  70. Xie, N.; Zhang, L.; Gao, W.; Huang, C.; Huber, P.E.; Zhou, X.; Li, C.; Shen, G.; Zou, B. NAD(+) metabolism: Pathophysiologic mechanisms and therapeutic potential. Signal Transduct. Target. Ther. 2020, 5, 227. [Google Scholar] [CrossRef]
  71. Mouchiroud, L.; Houtkooper, R.H.; Moullan, N.; Katsyuba, E.; Ryu, D.; Cantó, C.; Mottis, A.; Jo, Y.S.; Viswanathan, M.; Schoonjans, K.; et al. The NAD(+)/Sirtuin Pathway Modulates Longevity through Activation of Mitochondrial UPR and FOXO Signaling. Cell 2013, 154, 430–441. [Google Scholar] [CrossRef] [Green Version]
  72. Gomes, A.P.; Price, N.L.; Ling, A.J.; Moslehi, J.J.; Montgomery, M.K.; Rajman, L.; White, J.P.; Teodoro, J.S.; Wrann, C.D.; Hubbard, B.P.; et al. Declining NAD(+) induces a pseudohypoxic state disrupting nuclear-mitochondrial communication during aging. Cell 2013, 155, 1624–1638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Yoshino, J.; Mills, K.F.; Yoon, M.J.; Imai, S. Nicotinamide mononucleotide, a key NAD(+) intermediate, treats the pathophysiology of diet- and age-induced diabetes in mice. Cell Metab. 2011, 14, 528–536. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Senoo-Matsuda, N.; Yasuda, K.; Tsuda, M.; Ohkubo, T.; Yoshimura, S.; Nakazawa, H.; Hartman, P.S.; Ishii, N. A Defect in the Cytochrome b Large Subunit in Complex II Causes Both Superoxide Anion Overproduction and Abnormal Energy Metabolism in Caenorhabditis elegans *. J. Biol. Chem. 2001, 276, 41553–41558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Hosokawa, H.; Ishii, N.; Ishida, H.; Ichimori, K.; Nakazawa, H.; Suzuki, K. Rapid accumulation of fluorescent material with aging in an oxygen-sensitive mutant mev-1 of Caenorhabditis elegans. Mech. Ageing Dev. 1994, 74, 161–170. [Google Scholar] [CrossRef] [PubMed]
  76. Adachi, H.; Fujiwara, Y.; Ishii, N. Effects of oxygen on protein carbonyl and aging in Caenorhabditis elegans mutants with long (age-1) and short (mev-1) life spans. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 1998, 53, B240–B244. [Google Scholar] [CrossRef] [Green Version]
  77. Yoon, G.; Kim, H.J.; Yoon, Y.S.; Cho, H.; Lim, I.K.; Lee, J.H. Iron chelation-induced senescence-like growth arrest in hepatocyte cell lines: Association of transforming growth factor beta1 (TGF-beta1)-mediated p27Kip1 expression. Biochem. J. 2002, 366, 613–621. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Yoon, Y.S.; Byun, H.O.; Cho, H.; Kim, B.K.; Yoon, G. Complex II defect via down-regulation of iron-sulfur subunit induces mitochondrial dysfunction and cell cycle delay in iron chelation-induced senescence-associated growth arrest. J. Biol. Chem. 2003, 278, 51577–51586. [Google Scholar] [CrossRef] [Green Version]
  79. Byun, H.O.; Jung, H.J.; Kim, M.J.; Yoon, G. PKCδ phosphorylation is an upstream event of GSK3 inactivation-mediated ROS generation in TGF-β1-induced senescence. Free. Radic. Res. 2014, 48, 1100–1108. [Google Scholar] [CrossRef]
  80. Byun, H.O.; Jung, H.J.; Seo, Y.H.; Lee, Y.K.; Hwang, S.C.; Hwang, E.S.; Yoon, G. GSK3 inactivation is involved in mitochondrial complex IV defect in transforming growth factor (TGF) β1-induced senescence. Exp. Cell Res. 2012, 318, 1808–1819. [Google Scholar] [CrossRef]
  81. Yoon, Y.S.; Lee, J.H.; Hwang, S.C.; Choi, K.S.; Yoon, G. TGF beta1 induces prolonged mitochondrial ROS generation through decreased complex IV activity with senescent arrest in Mv1Lu cells. Oncogene 2005, 24, 1895–1903. [Google Scholar] [CrossRef] [Green Version]
  82. Sakellariou, G.K.; Pye, D.; Vasilaki, A.; Zibrik, L.; Palomero, J.; Kabayo, T.; McArdle, F.; Van Remmen, H.; Richardson, A.; Tidball, J.G.; et al. Role of superoxide-nitric oxide interactions in the accelerated age-related loss of muscle mass in mice lacking Cu, Zn superoxide dismutase. Aging Cell 2011, 10, 749–760. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Llanos-González, E.; Henares-Chavarino, Á.A.; Pedrero-Prieto, C.M.; García-Carpintero, S.; Frontiñán-Rubio, J.; Sancho-Bielsa, F.J.; Alcain, F.J.; Peinado, J.R.; Rabanal-Ruíz, Y.; Durán-Prado, M. Interplay Between Mitochondrial Oxidative Disorders and Proteostasis in Alzheimer’s Disease. Front. Neurosci. 2019, 13, 1444. [Google Scholar] [CrossRef] [PubMed]
  84. Jensen, M.B.; Jasper, H. Mitochondrial Proteostasis in the Control of Aging and Longevity. Cell Metab. 2014, 20, 214–225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Bulteau, A.L.; Mena, N.P.; Auchère, F.; Lee, I.; Prigent, A.; Lobsiger, C.S.; Camadro, J.M.; Hirsch, E.C. Dysfunction of mitochondrial Lon protease and identification of oxidized protein in mouse brain following exposure to MPTP: Implications for Parkinson disease. Free. Radic. Biol. Med. 2017, 108, 236–246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Alikhani, N.; Guo, L.; Yan, S.; Du, H.; Pinho, C.M.; Chen, J.X.; Glaser, E.; Yan, S.S. Decreased proteolytic activity of the mitochondrial amyloid-β degrading enzyme, PreP peptidasome, in Alzheimer’s disease brain mitochondria. J. Alzheimer’s Dis. JAD 2011, 27, 75–87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Takenaka, Y.; Inoue, I.; Nakano, T.; Ikeda, M.; Kakinuma, Y. Prolonged disturbance of proteostasis induces cellular senescence via temporal mitochondrial dysfunction and subsequent mitochondrial accumulation in human fibroblasts. FEBS J. 2022, 289, 1650–1667. [Google Scholar] [CrossRef] [PubMed]
  88. Paez, H.G.; Pitzer, C.R.; Alway, S.E. Age-Related Dysfunction in Proteostasis and Cellular Quality Control in the Development of Sarcopenia. Cells 2023, 12, 249. [Google Scholar] [CrossRef]
  89. Ichimura, H.; Parthasarathi, K.; Quadri, S.; Issekutz, A.C.; Bhattacharya, J. Mechano-oxidative coupling by mitochondria induces proinflammatory responses in lung venular capillaries. J. Clin. Investig. 2003, 111, 691–699. [Google Scholar] [CrossRef] [Green Version]
  90. Naik, E.; Dixit, V.M. Mitochondrial reactive oxygen species drive proinflammatory cytokine production. J. Exp. Med. 2011, 208, 417–420. [Google Scholar] [CrossRef] [Green Version]
  91. Nelson, G.; Kucheryavenko, O.; Wordsworth, J.; von Zglinicki, T. The senescent bystander effect is caused by ROS-activated NF-κB signalling. Mech. Ageing Dev. 2018, 170, 30–36. [Google Scholar] [CrossRef]
  92. Leyane, T.S.; Jere, S.W.; Houreld, N.N. Oxidative Stress in Ageing and Chronic Degenerative Pathologies: Molecular Mechanisms Involved in Counteracting Oxidative Stress and Chronic Inflammation. Int. J. Mol. Sci. 2022, 23, 7273. [Google Scholar] [CrossRef]
  93. Correia-Melo, C.; Marques, F.D.; Anderson, R.; Hewitt, G.; Hewitt, R.; Cole, J.; Carroll, B.M.; Miwa, S.; Birch, J.; Merz, A.; et al. Mitochondria are required for pro-ageing features of the senescent phenotype. EMBO J. 2016, 35, 724–742. [Google Scholar] [CrossRef]
  94. Taanman, J.-W. The mitochondrial genome: Structure, transcription, translation and replication. Biochim. Biophys. Acta (BBA) Bioenerg. 1999, 1410, 103–123. [Google Scholar] [CrossRef] [Green Version]
  95. Mikhed, Y.; Daiber, A.; Steven, S. Mitochondrial Oxidative Stress, Mitochondrial DNA Damage and Their Role in Age-Related Vascular Dysfunction. Int. J. Mol. Sci. 2015, 16, 15918–15953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Simmons, R.A.; Suponitsky-Kroyter, I.; Selak, M.A. Progressive accumulation of mitochondrial DNA mutations and decline in mitochondrial function lead to beta-cell failure. J. Biol. Chem. 2005, 280, 28785–28791. [Google Scholar] [CrossRef] [Green Version]
  97. Van Houten, B.; Woshner, V.; Santos, J.H. Role of mitochondrial DNA in toxic responses to oxidative stress. DNA Repair 2006, 5, 145–152. [Google Scholar] [CrossRef]
  98. Aliev, G.; Seyidova, D.; Lamb, B.T.; Obrenovich, M.E.; Siedlak, S.L.; Vinters, H.V.; Friedland, R.P.; Siedlak, S.L.; Vinters, H.V.; Friedland, R.P.; et al. Mitochondria and vascular lesions as a central target for the development of Alzheimer’s disease and Alzheimer disease-like pathology in transgenic mice. Neurol. Res. 2003, 25, 665–674. [Google Scholar] [CrossRef] [PubMed]
  99. Swerdlow, R.H.; Parks, J.K.; Cassarino, D.S.; Maguire, D.J.; Maguire, R.S.; Bennett, J.P., Jr.; Davis, R.E.; Parker, W.D., Jr. Cybrids in Alzheimer’s disease: A cellular model of the disease? Neurology 1997, 49, 918–925. [Google Scholar] [CrossRef] [PubMed]
  100. Kukreja, L.; Kujoth, G.C.; Prolla, T.A.; Van Leuven, F.; Vassar, R. Increased mtDNA mutations with aging promotes amyloid accumulation and brain atrophy in the APP/Ld transgenic mouse model of Alzheimer’s disease. Mol. Neurodegener. 2014, 9, 16. [Google Scholar] [CrossRef] [Green Version]
  101. Barja, G.; Herrero, A. Oxidative damage to mitochondrial DNA is inversely related to maximum life span in the heart and brain of mammals. FASEB J. 2000, 14, 312–318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Onur, S.; Niklowitz, P.; Fischer, A.; Metges, C.C.; Grune, T.; Menke, T.; Rimbach, G.; Döring, F. A comparative study into alterations of coenzyme Q redox status in ageing pigs, mice, and worms. BioFactors 2014, 40, 346–354. [Google Scholar] [CrossRef] [PubMed]
  103. Lass, A.; Kwong, L.; Sohal, R.S. Mitochondrial coenzyme Q content and aging. BioFactors 1999, 9, 199–205. [Google Scholar] [CrossRef] [PubMed]
  104. Kamzalov, S.; Sohal, R.S. Effect of age and caloric restriction on coenzyme Q and alpha-tocopherol levels in the rat. Exp. Gerontol. 2004, 39, 1199–1205. [Google Scholar] [CrossRef] [PubMed]
  105. De la Mata, M.; Cotán, D.; Oropesa-Ávila, M.; Garrido-Maraver, J.; Cordero, M.D.; Villanueva Paz, M.; Delgado Pavón, A.; Alcocer-Gómez, E.; de Lavera, I.; Ybot-González, P.; et al. Pharmacological Chaperones and Coenzyme Q10 Treatment Improves Mutant β-Glucocerebrosidase Activity and Mitochondrial Function in Neuronopathic Forms of Gaucher Disease. Sci. Rep. 2015, 5, 10903. [Google Scholar] [CrossRef] [Green Version]
  106. Wani, W.Y.; Gudup, S.; Sunkaria, A.; Bal, A.; Singh, P.P.; Kandimalla, R.J.; Sharma, D.R.; Gill, K.D. Protective efficacy of mitochondrial targeted antioxidant MitoQ against dichlorvos induced oxidative stress and cell death in rat brain. Neuropharmacology 2011, 61, 1193–1201. [Google Scholar] [CrossRef]
  107. Tian, G.; Sawashita, J.; Kubo, H.; Nishio, S.Y.; Hashimoto, S.; Suzuki, N.; Yoshimura, H.; Tsuruoka, M.; Wang, Y.; Liu, Y.; et al. Ubiquinol-10 supplementation activates mitochondria functions to decelerate senescence in senescence-accelerated mice. Antioxid. Redox Signal. 2014, 20, 2606–2620. [Google Scholar] [CrossRef]
  108. Xie, J.; Lin, J.; Wei, M.; Teng, Y.; He, Q.; Yang, G.; Yang, X. Sustained Akt signaling in articular chondrocytes causes osteoarthritis via oxidative stress-induced senescence in mice. Bone Res. 2019, 7, 23. [Google Scholar] [CrossRef] [Green Version]
  109. Reddy, T.P.; Manczak, M.; Calkins, M.J.; Mao, P.; Reddy, A.P.; Shirendeb, U.; Park, B.; Reddy, P.H. Toxicity of neurons treated with herbicides and neuroprotection by mitochondria-targeted antioxidant SS31. Int. J. Environ. Res. Public Health 2011, 8, 203–221. [Google Scholar] [CrossRef] [Green Version]
  110. Yang, L.; Zhao, K.; Calingasan, N.Y.; Luo, G.; Szeto, H.H.; Beal, M.F. Mitochondria targeted peptides protect against 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine neurotoxicity. Antioxid. Redox Signal. 2009, 11, 2095–2104. [Google Scholar] [CrossRef] [Green Version]
  111. Calkins, M.J.; Manczak, M.; Reddy, P.H. Mitochondria-Targeted Antioxidant SS31 Prevents Amyloid Beta-Induced Mitochondrial Abnormalities and Synaptic Degeneration in Alzheimer’s Disease. Pharmaceuticals 2012, 5, 1103–1119. [Google Scholar] [CrossRef] [Green Version]
  112. Cho, S.; Szeto, H.H.; Kim, E.; Kim, H.; Tolhurst, A.T.; Pinto, J.T. A novel cell-permeable antioxidant peptide, SS31, attenuates ischemic brain injury by down-regulating CD36. J. Biol. Chem. 2007, 282, 4634–4642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Das, K.; Roychoudhury, A. Reactive oxygen species (ROS) and response of antioxidants as ROS-scavengers during environmental stress in plants. Front. Environ. Sci. 2014, 2, 53. [Google Scholar] [CrossRef] [Green Version]
  114. Schriner, S.E.; Linford, N.J. Extension of mouse lifespan by overexpression of catalase. Age 2006, 28, 209–218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Ashrafi, G.; Schwarz, T.L. The pathways of mitophagy for quality control and clearance of mitochondria. Cell Death Differ. 2013, 20, 31–42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Ma, S.; Zhang, X.; Zheng, L.; Li, Z.; Zhao, X.; Lai, W.; Shen, H.; Lv, J.; Yang, G.; Wang, Q.; et al. Peroxiredoxin 6 Is a Crucial Factor in the Initial Step of Mitochondrial Clearance and Is Upstream of the PINK1-Parkin Pathway. Antioxid. Redox Signal. 2016, 24, 486–501. [Google Scholar] [CrossRef]
  117. Leduc-Gaudet, J.-P.; Reynaud, O.; Hussain, S.N.; Gouspillou, G. Parkin overexpression protects from ageing-related loss of muscle mass and strength. J. Physiol. 2019, 597, 1975–1991. [Google Scholar] [CrossRef]
  118. Gouspillou, G.; Godin, R.; Piquereau, J.; Picard, M.; Mofarrahi, M.; Mathew, J.; Purves-Smith, F.M.; Sgarioto, N.; Hepple, R.T.; Burelle, Y.; et al. Protective role of Parkin in skeletal muscle contractile and mitochondrial function. J. Physiol. 2018, 596, 2565–2579. [Google Scholar] [CrossRef]
  119. Liu, H.; Zang, C.; Yuan, F.; Ju, C.; Shang, M.; Ning, J.; Yang, Y.; Ma, J.; Li, G.; Bao, X.; et al. The role of FUNDC1 in mitophagy, mitochondrial dynamics and human diseases. Biochem. Pharmacol. 2022, 197, 114891. [Google Scholar] [CrossRef]
  120. Schmid, E.T.; Pyo, J.-H.; Walker, D.W. Neuronal induction of BNIP3-mediated mitophagy slows systemic aging in Drosophila. Nat. Aging 2022, 2, 494–507. [Google Scholar] [CrossRef]
  121. Kim, Y.J.; Choo, O.-S.; Lee, J.-S.; Jang, J.H.; Woo, H.G.; Choung, Y.-H. BCL2 Interacting Protein 3-like/NIX-mediated Mitophagy Plays an Important Role in the Process of Age-related Hearing Loss. Neuroscience 2021, 455, 39–51. [Google Scholar] [CrossRef]
  122. LaRocca, T.J.; Hearon, C.M., Jr.; Henson, G.D.; Seals, D.R. Mitochondrial quality control and age-associated arterial stiffening. Exp. Gerontol. 2014, 58, 78–82. [Google Scholar] [CrossRef] [Green Version]
  123. Tam, Z.Y.; Gruber, J.; Ng, L.F.; Halliwell, B.; Gunawan, R. Effects of lithium on age-related decline in mitochondrial turnover and function in Caenorhabditis elegans. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 2014, 69, 810–820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Lotfi, P.; Tse, D.Y.; Di Ronza, A.; Seymour, M.L.; Martano, G.; Cooper, J.D.; Pereira, F.A.; Passafaro, M.; Wu, S.M.; Sardiello, M. Trehalose reduces retinal degeneration, neuroinflammation and storage burden caused by a lysosomal hydrolase deficiency. Autophagy 2018, 14, 1419–1434. [Google Scholar] [CrossRef] [Green Version]
  125. Yang, B.; Dan, X.; Hou, Y.; Lee, J.-H.; Wechter, N.; Krishnamurthy, S.; Kimura, R.; Babbar, M.; Demarest, T.; McDevitt, R.; et al. NAD+ supplementation prevents STING-induced senescence in ataxia telangiectasia by improving mitophagy. Aging Cell 2021, 20, e13329. [Google Scholar] [CrossRef] [PubMed]
  126. Fang, E.F.; Scheibye-Knudsen, M.; Brace, L.E.; Kassahun, H.; SenGupta, T.; Nilsen, H.; Mitchell, J.R.; Croteau, D.L.; Bohr, V.A. Defective mitophagy in XPA via PARP-1 hyperactivation and NAD(+)/SIRT1 reduction. Cell 2014, 157, 882–896. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Kang, H.T.; Park, J.T.; Choi, K.; Kim, Y.; Choi, H.J.C.; Jung, C.W.; Lee, Y.S.; Park, S.C. Chemical screening identifies ATM as a target for alleviating senescence. Nat. Chem. Biol. 2017, 13, 616–623. [Google Scholar] [CrossRef]
  128. Kuk, M.U.; Kim, J.W.; Lee, Y.S.; Cho, K.A.; Park, J.T.; Park, S.C. Alleviation of Senescence via ATM Inhibition in Accelerated Aging Models. Mol. Cells 2019, 42, 210–217. [Google Scholar] [CrossRef]
  129. Kang, H.T.; Park, J.T.; Choi, K.; Choi, H.J.C.; Jung, C.W.; Kim, G.R.; Lee, Y.S.; Park, S.C. Chemical screening identifies ROCK as a target for recovering mitochondrial function in Hutchinson-Gilford progeria syndrome. Aging Cell 2017, 16, 541–550. [Google Scholar] [CrossRef]
  130. Park, J.T.; Kang, H.T.; Park, C.H.; Lee, Y.S.; Cho, K.A.; Park, S.C. A crucial role of ROCK for alleviation of senescence-associated phenotype. Exp. Gerontol. 2018, 106, 8–15. [Google Scholar] [CrossRef]
  131. Olgun, A.; Akman, S.; Serdar, M.A.; Kutluay, T. Oxidative phosphorylation enzyme complexes in caloric restriction. Exp. Gerontol. 2002, 37, 639–645. [Google Scholar] [CrossRef]
  132. Masoro, E.J. Retardation of aging processes by food restriction: An experimental tool. Am. J. Clin. Nutr. 1992, 55, 1250S–1252S. [Google Scholar] [CrossRef]
  133. Sayer, A.A.; Cooper, C. Undernutrition and Aging. Gerontology 1997, 43, 203–205. [Google Scholar] [CrossRef] [PubMed]
  134. Valenti, D.; De Rasmo, D.; Signorile, A.; Rossi, L.; de Bari, L.; Scala, I.; Granese, B.; Papa, S.; Vacca, R.A. Epigallocatechin-3-gallate prevents oxidative phosphorylation deficit and promotes mitochondrial biogenesis in human cells from subjects with Down’s syndrome. Biochim. Biophys. Acta 2013, 1832, 542–552. [Google Scholar] [CrossRef] [Green Version]
  135. Sharma, R.; Sharma, A.; Kumari, A.; Kulurkar, P.M.; Raj, R.; Gulati, A.; Padwad, Y.S. Consumption of green tea epigallocatechin-3-gallate enhances systemic immune response, antioxidative capacity and HPA axis functions in aged male swiss albino mice. Biogerontology 2017, 18, 367–382. [Google Scholar] [CrossRef] [PubMed]
  136. Niu, Y.; Na, L.; Feng, R.; Gong, L.; Zhao, Y.; Li, Q.; Li, Y.; Sun, C. The phytochemical, EGCG, extends lifespan by reducing liver and kidney function damage and improving age-associated inflammation and oxidative stress in healthy rats. Aging Cell 2013, 12, 1041–1049. [Google Scholar] [CrossRef] [PubMed]
  137. Brown, M.K.; Evans, J.L.; Luo, Y. Beneficial effects of natural antioxidants EGCG and alpha-lipoic acid on life span and age-dependent behavioral declines in Caenorhabditis elegans. Pharmacol. Biochem. Behav. 2006, 85, 620–628. [Google Scholar] [CrossRef]
  138. Zhang, L.; Jie, G.; Zhang, J.; Zhao, B. Significant longevity-extending effects of EGCG on Caenorhabditis elegans under stress. Free. Radic. Biol. Med. 2009, 46, 414–421. [Google Scholar] [CrossRef]
  139. Marcu, R.; Wiczer, B.M.; Neeley, C.K.; Hawkins, B.J. Mitochondrial matrix Ca2+ accumulation regulates cytosolic NAD;+/NADH metabolism, protein acetylation, and sirtuin expression. Mol. Cell. Biol. 2014, 34, 2890–2902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Müller, M.; Ahumada-Castro, U.; Sanhueza, M.; Gonzalez-Billault, C.; Court, F.A.; Cárdenas, C. Mitochondria and Calcium Regulation as Basis of Neurodegeneration Associated with Aging. Front. Neurosci. 2018, 12, 470. [Google Scholar] [CrossRef]
  141. Gherardi, G.; Monticelli, H.; Rizzuto, R.; Mammucari, C. The Mitochondrial Ca2+ Uptake and the Fine-Tuning of Aerobic Metabolism. Front. Physiol. 2020, 11, 554904. [Google Scholar] [CrossRef]
  142. Jahangir, A.; Sagar, S.; Terzic, A. Aging and cardioprotection. J. Appl. Physiol. 2007, 103, 2120–2128. [Google Scholar] [CrossRef] [PubMed]
  143. Kuznetsov, A.V.; Javadov, S.; Margreiter, R.; Grimm, M.; Hagenbuchner, J.; Ausserlechner, M.J. The Role of Mitochondria in the Mechanisms of Cardiac Ischemia-Reperfusion Injury. Antioxidants 2019, 8, 454. [Google Scholar] [CrossRef] [Green Version]
  144. Woods, J.J.; Nemani, N.; Shanmughapriya, S.; Kumar, A.; Zhang, M.; Nathan, S.R.; Thomas, M.; Carvalho, E.; Ramachandran, K.; Srikantan, S.; et al. A Selective and Cell-Permeable Mitochondrial Calcium Uniporter (MCU) Inhibitor Preserves Mitochondrial Bioenergetics after Hypoxia/Reoxygenation Injury. ACS Cent. Sci. 2019, 5, 153–166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. de Jesús García-Rivas, G.; Guerrero-Hernández, A.; Guerrero-Serna, G.; Rodríguez-Zavala, J.S.; Zazueta, C. Inhibition of the mitochondrial calcium uniporter by the oxo-bridged dinuclear ruthenium amine complex (Ru360) prevents from irreversible injury in postischemic rat heart. FEBS J. 2005, 272, 3477–3488. [Google Scholar] [CrossRef] [PubMed]
  146. Sripetchwandee, J.; Sanit, J.; Chattipakorn, N.; Chattipakorn, S.C. Mitochondrial calcium uniporter blocker effectively prevents brain mitochondrial dysfunction caused by iron overload. Life Sci. 2013, 92, 298–304. [Google Scholar] [CrossRef] [PubMed]
  147. Pan, L.; Huang, B.J.; Ma, X.E.; Wang, S.Y.; Feng, J.; Lv, F.; Liu, Y.; Liu, Y.; Li, C.M.; Liang, D.D.; et al. MiR-25 protects cardiomyocytes against oxidative damage by targeting the mitochondrial calcium uniporter. Int. J. Mol. Sci. 2015, 16, 5420–5433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. St-Pierre, J.; Drori, S.; Uldry, M.; Silvaggi, J.; Rhee, J.; Jager, S.; Handschin, C.; Zheng, K.; Lin, J.; Yang, W.; et al. Suppression of reactive oxygen species and neurodegeneration by the PGC-1 transcriptional coactivators. Cell 2006, 127, 397–408. [Google Scholar] [CrossRef] [Green Version]
  149. Cao, C.; Lu, S.; Kivlin, R.; Wallin, B.; Card, E.; Bagdasarian, A.; Tamakloe, T.; Wang, W.J.; Song, X.; Chu, W.M.; et al. SIRT1 confers protection against UVB- and H2O2-induced cell death via modulation of p53 and JNK in cultured skin keratinocytes. J. Cell. Mol. Med. 2009, 13, 3632–3643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Merksamer, P.I.; Liu, Y.; He, W.; Hirschey, M.D.; Chen, D.; Verdin, E. The sirtuins, oxidative stress and aging: An emerging link. Aging 2013, 5, 144–150. [Google Scholar] [CrossRef] [Green Version]
  151. Lozada-Delgado, J.G.; Torres-Ramos, C.A.; Ayala-Peña, S. Chapter 4—Aging, oxidative stress, mitochondrial dysfunction, and the liver. In Aging, 2nd ed.; Preedy, V.R., Patel, V.B., Eds.; Academic Press: Cambridge, MA, USA, 2020; pp. 37–46. [Google Scholar]
  152. Cui, H.; Kong, Y.; Zhang, H. Oxidative stress, mitochondrial dysfunction, and aging. J. Signal Transduct. 2012, 2012, 646354. [Google Scholar] [CrossRef] [Green Version]
  153. Dai, D.F.; Chiao, Y.A.; Marcinek, D.J.; Szeto, H.H.; Rabinovitch, P.S. Mitochondrial oxidative stress in aging and healthspan. Longev. Healthspan 2014, 3, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Conrad, C.; Lymp, J.; Thompson, V.; Dunn, C.; Davies, Z.; Chatfield, B.; Nichols, D.; Clancy, J.; Vender, R.; Egan, M.E.; et al. Long-term treatment with oral N-acetylcysteine: Affects lung function but not sputum inflammation in cystic fibrosis subjects. A phase II randomized placebo-controlled trial. J. Cyst. Fibros. Off. J. Eur. Cyst. Fibros. Soc. 2015, 14, 219–227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Xu, R.; Tao, A.; Bai, Y.; Deng, Y.; Chen, G. Effectiveness of N-Acetylcysteine for the Prevention of Contrast-Induced Nephropathy: A Systematic Review and Meta-Analysis of Randomized Controlled Trials. J. Am. Heart Assoc. 2016, 5, e003968. [Google Scholar] [CrossRef] [Green Version]
  156. Jiang, Q.; Yin, J.; Chen, J.; Ma, X.; Wu, M.; Liu, G.; Yao, K.; Tan, B.; Yin, Y. Mitochondria-Targeted Antioxidants: A Step towards Disease Treatment. Oxidative Med. Cell. Longev. 2020, 2020, 8837893. [Google Scholar] [CrossRef]
  157. Fields, M.; Marcuzzi, A.; Gonelli, A.; Celeghini, C.; Maximova, N.; Rimondi, E. Mitochondria-Targeted Antioxidants, an Innovative Class of Antioxidant Compounds for Neurodegenerative Diseases: Perspectives and Limitations. Int. J. Mol. Sci. 2023, 24, 3739. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Mitochondrial alterations during the process of senescence and aging: (A) Mitochondria maintain their morphology, quality, and function through mitochondrial dynamics consisting of fusion and fission. Proteins involved in mitochondrial fusion: mitofusin 1 (Mfn1), mitofusin 2 (Mfn2), and OPA1. Proteins involved in mitochondrial fission: mitochondrial dynamics protein of 49 kDa and 51 kDa (MiD49 and MiD51, respectively), mitochondrial fission factor (Mff), mitochondrial fission 1 protein (FIS1), and dynamin-related protein 1 (Drp1). (B) Senescent and H2O2-induced senescent cells express low levels of FIS1, forming large mitochondria with highly interconnected network structures. ROS: reactive oxygen species. The lightning bolt represents senescence-associated stress.
Figure 1. Mitochondrial alterations during the process of senescence and aging: (A) Mitochondria maintain their morphology, quality, and function through mitochondrial dynamics consisting of fusion and fission. Proteins involved in mitochondrial fusion: mitofusin 1 (Mfn1), mitofusin 2 (Mfn2), and OPA1. Proteins involved in mitochondrial fission: mitochondrial dynamics protein of 49 kDa and 51 kDa (MiD49 and MiD51, respectively), mitochondrial fission factor (Mff), mitochondrial fission 1 protein (FIS1), and dynamin-related protein 1 (Drp1). (B) Senescent and H2O2-induced senescent cells express low levels of FIS1, forming large mitochondria with highly interconnected network structures. ROS: reactive oxygen species. The lightning bolt represents senescence-associated stress.
Antioxidants 12 00934 g001
Figure 2. Fundamental mechanism of mitochondrial Ca2+ homeostasis: (A) Mitochondrial homeostasis is primarily regulated by mitochondrial Ca2+ concentration. Mitochondrial Ca2+ concentration is controlled by channels in the mitochondria and in the endoplasmic reticulum (ER). VDAC: voltage-dependent anion channels, MCU: mitochondrial calcium uniporter, HCX: H+/Ca2+ exchanger, NCLX: Na+/Ca2+ exchanger, IP3R: inositol 1,4,5-trisphosphate receptor. Orange dots represent Ca2+. (B) Senescence causes efflux of Ca2+ from the IP3R. Ca2+ then starts to be transported to the mitochondria through the VDAC/MCU channel. Mitochondrial Ca2+ overload induces mitochondrial ROS generation and senescence; the detailed mechanism will be further discussed in Section 3. Orange dots represent Ca2+. Lightning bolts represent senescence-associated stress. Pink arrows represent that large amounts of Ca2+ are being transported to the mitochondria through the VDAC/MCU channel.
Figure 2. Fundamental mechanism of mitochondrial Ca2+ homeostasis: (A) Mitochondrial homeostasis is primarily regulated by mitochondrial Ca2+ concentration. Mitochondrial Ca2+ concentration is controlled by channels in the mitochondria and in the endoplasmic reticulum (ER). VDAC: voltage-dependent anion channels, MCU: mitochondrial calcium uniporter, HCX: H+/Ca2+ exchanger, NCLX: Na+/Ca2+ exchanger, IP3R: inositol 1,4,5-trisphosphate receptor. Orange dots represent Ca2+. (B) Senescence causes efflux of Ca2+ from the IP3R. Ca2+ then starts to be transported to the mitochondria through the VDAC/MCU channel. Mitochondrial Ca2+ overload induces mitochondrial ROS generation and senescence; the detailed mechanism will be further discussed in Section 3. Orange dots represent Ca2+. Lightning bolts represent senescence-associated stress. Pink arrows represent that large amounts of Ca2+ are being transported to the mitochondria through the VDAC/MCU channel.
Antioxidants 12 00934 g002
Figure 3. Defective mitochondria are a major cause of mitochondrial ROS generation: (A) Complex I damaged by oxidative stress inefficiently transports electrons and subsequently increases electron leakage to oxygen, generating superoxide anions (O2). Superoxide dismutase (SOD) converts superoxide anions (O2) to hydrogen peroxide (H2O2). Then, hydrogen peroxide (H2O2) is partially reduced by the Fenton reaction and converted to a more harmful free radical, the hydroxyl radical (OH). e: electron. (B) Mitochondrial Ca2+ overload stimulates the mitochondrial permeability transition (PT) and opens PT pores (mPTP), allowing ions and other solutes to move freely. Enlarged mitochondria readily lose cytochrome c that is loosely bound to the IMM. Loss of cytochrome c impedes electron transport from complex I to IV, resulting in increased electron leakage from ETC. The leaked electrons react with oxygen to generate large amounts of mitochondrial ROS. Purple dots represent ions and solutes. Purple arrows represent that large amounts of ions and solutes are being transported to mitochondria via mPTP.
Figure 3. Defective mitochondria are a major cause of mitochondrial ROS generation: (A) Complex I damaged by oxidative stress inefficiently transports electrons and subsequently increases electron leakage to oxygen, generating superoxide anions (O2). Superoxide dismutase (SOD) converts superoxide anions (O2) to hydrogen peroxide (H2O2). Then, hydrogen peroxide (H2O2) is partially reduced by the Fenton reaction and converted to a more harmful free radical, the hydroxyl radical (OH). e: electron. (B) Mitochondrial Ca2+ overload stimulates the mitochondrial permeability transition (PT) and opens PT pores (mPTP), allowing ions and other solutes to move freely. Enlarged mitochondria readily lose cytochrome c that is loosely bound to the IMM. Loss of cytochrome c impedes electron transport from complex I to IV, resulting in increased electron leakage from ETC. The leaked electrons react with oxygen to generate large amounts of mitochondrial ROS. Purple dots represent ions and solutes. Purple arrows represent that large amounts of ions and solutes are being transported to mitochondria via mPTP.
Antioxidants 12 00934 g003
Figure 4. Vicious feedback loop between mitochondrial oxidative stress and senescence/aging. (A) Accumulating oxidative damage over time activates the p53 and Rb (retinoblastoma protein) pathways, resulting in persistent cell cycle arrest and senescence. Mitochondrial ROS also activates polyADP-ribose polymerase 1, which increases NAD+ consumption, thereby reducing NAD+ levels. A decrease in NAD+ levels leads to cellular dysfunction and aggravates age-related pathologies. (B) Mitochondrial proteostasis consists of chaperones that promote protein folding and proteases that clearly damage mitochondrial proteins. The prolonged disturbance of mitochondrial proteostasis by intracellular and mitochondrial ROS causes the accumulation of misfolded or aggregated mitochondrial proteins and triggers cellular senescence. (C) Mitochondrial oxidative stress activates the secretion of inflammatory cytokines and proinflammatory senescence-associated secretory phenotype (SASP). Different colored dots represent inflammatory cytokines and proinflammatory SASP. (D) Mitochondrial oxidative stress directly damages mtDNA, as mitochondrial ROS is produced in the mtDNA-containing mitochondrial matrix.
Figure 4. Vicious feedback loop between mitochondrial oxidative stress and senescence/aging. (A) Accumulating oxidative damage over time activates the p53 and Rb (retinoblastoma protein) pathways, resulting in persistent cell cycle arrest and senescence. Mitochondrial ROS also activates polyADP-ribose polymerase 1, which increases NAD+ consumption, thereby reducing NAD+ levels. A decrease in NAD+ levels leads to cellular dysfunction and aggravates age-related pathologies. (B) Mitochondrial proteostasis consists of chaperones that promote protein folding and proteases that clearly damage mitochondrial proteins. The prolonged disturbance of mitochondrial proteostasis by intracellular and mitochondrial ROS causes the accumulation of misfolded or aggregated mitochondrial proteins and triggers cellular senescence. (C) Mitochondrial oxidative stress activates the secretion of inflammatory cytokines and proinflammatory senescence-associated secretory phenotype (SASP). Different colored dots represent inflammatory cytokines and proinflammatory SASP. (D) Mitochondrial oxidative stress directly damages mtDNA, as mitochondrial ROS is produced in the mtDNA-containing mitochondrial matrix.
Antioxidants 12 00934 g004
Figure 5. Targeting mitochondrial oxidative stress as a therapeutic strategy for aging and age-related diseases: (A) Senescent cells show a lack of coenzyme Q (CoQ), which transports electrons from complex II to III (green CoQ indicates CoQ deficiency). Senescent cells deficient in CoQ exhibit electron leakage as electron transport in ETC is hindered. Premature electron leakage from ETC combines with oxygen to generate mitochondrial ROS. By contrast, mitochondria-targeted CoQ (MitoQ) treatment lowers mitochondrial ROS production, consequently resulting in a significant reduction in senescence-associated symptoms (pink CoQ indicates high levels of CoQ); e: electron. (B) Mitochondria damaged by high levels of mitochondrial ROS recruit PTEN-induced putative kinase protein 1 (PINK1) and ubiquitin ligase PARKIN. PINK1 activates PARKIN recruitment to the OMM for ubiquitylation of OMM proteins in damaged mitochondria. The creation of polyubiquitin chains in OMM proteins recruits autophagy receptor proteins that interact with LC3 (microtubule-associated protein 1A/1B light chain 3) in the phagophore. Then, damaged mitochondria are eliminated by the subsequent formation of the phagophores, autophagosomes, and autolysosomes. (C) Reduction in mitochondrial oxidative damage by directly targeting genes that regulate the activity of ETC components. ROCK activation phosphorylates Rac1b, making it easier for Rac1b to interact with cytochrome c. This interaction steals electrons from cytochrome c and causes a partial reduction in oxygen, triggering mitochondrial ROS production. Inhibition of ROCK activity prevents Rac1b from intercepting electrons from cytochrome c, enabling electron transport from complex III to IV. Then, efficient electron transport facilitates complex IV activity, thereby reducing mitochondrial ROS production; e: electron. (D) Inhibition of MCU by ruthenium 360 (Ru360) reduces the percentage of mitochondria exhibiting Ca2+ overload and subsequently reduces the production of mitochondrial ROS. MicroRNA-mediated silencing of MCU shields cardiomyocytes from mitochondrial oxidative stress.
Figure 5. Targeting mitochondrial oxidative stress as a therapeutic strategy for aging and age-related diseases: (A) Senescent cells show a lack of coenzyme Q (CoQ), which transports electrons from complex II to III (green CoQ indicates CoQ deficiency). Senescent cells deficient in CoQ exhibit electron leakage as electron transport in ETC is hindered. Premature electron leakage from ETC combines with oxygen to generate mitochondrial ROS. By contrast, mitochondria-targeted CoQ (MitoQ) treatment lowers mitochondrial ROS production, consequently resulting in a significant reduction in senescence-associated symptoms (pink CoQ indicates high levels of CoQ); e: electron. (B) Mitochondria damaged by high levels of mitochondrial ROS recruit PTEN-induced putative kinase protein 1 (PINK1) and ubiquitin ligase PARKIN. PINK1 activates PARKIN recruitment to the OMM for ubiquitylation of OMM proteins in damaged mitochondria. The creation of polyubiquitin chains in OMM proteins recruits autophagy receptor proteins that interact with LC3 (microtubule-associated protein 1A/1B light chain 3) in the phagophore. Then, damaged mitochondria are eliminated by the subsequent formation of the phagophores, autophagosomes, and autolysosomes. (C) Reduction in mitochondrial oxidative damage by directly targeting genes that regulate the activity of ETC components. ROCK activation phosphorylates Rac1b, making it easier for Rac1b to interact with cytochrome c. This interaction steals electrons from cytochrome c and causes a partial reduction in oxygen, triggering mitochondrial ROS production. Inhibition of ROCK activity prevents Rac1b from intercepting electrons from cytochrome c, enabling electron transport from complex III to IV. Then, efficient electron transport facilitates complex IV activity, thereby reducing mitochondrial ROS production; e: electron. (D) Inhibition of MCU by ruthenium 360 (Ru360) reduces the percentage of mitochondria exhibiting Ca2+ overload and subsequently reduces the production of mitochondrial ROS. MicroRNA-mediated silencing of MCU shields cardiomyocytes from mitochondrial oxidative stress.
Antioxidants 12 00934 g005
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

Lee, Y.H.; Kuk, M.U.; So, M.K.; Song, E.S.; Lee, H.; Ahn, S.K.; Kwon, H.W.; Park, J.T.; Park, S.C. Targeting Mitochondrial Oxidative Stress as a Strategy to Treat Aging and Age-Related Diseases. Antioxidants 2023, 12, 934. https://doi.org/10.3390/antiox12040934

AMA Style

Lee YH, Kuk MU, So MK, Song ES, Lee H, Ahn SK, Kwon HW, Park JT, Park SC. Targeting Mitochondrial Oxidative Stress as a Strategy to Treat Aging and Age-Related Diseases. Antioxidants. 2023; 12(4):934. https://doi.org/10.3390/antiox12040934

Chicago/Turabian Style

Lee, Yun Haeng, Myeong Uk Kuk, Moon Kyoung So, Eun Seon Song, Haneur Lee, Soon Kil Ahn, Hyung Wook Kwon, Joon Tae Park, and Sang Chul Park. 2023. "Targeting Mitochondrial Oxidative Stress as a Strategy to Treat Aging and Age-Related Diseases" Antioxidants 12, no. 4: 934. https://doi.org/10.3390/antiox12040934

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