Next Article in Journal
Arginine-Coated Nanoglobules for the Nasal Delivery of Insulin
Next Article in Special Issue
The Mitochondrion: A Promising Target for Kidney Disease
Previous Article in Journal
Delivery of Therapeutic Biopolymers Employing Silica-Based Nanosystems
Previous Article in Special Issue
Identification of Autophagy as a Functional Target Suitable for the Pharmacological Treatment of Mitochondrial Membrane Protein-Associated Neurodegeneration (MPAN) In Vitro
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Targeting Mitochondria to Control Ageing and Senescence

1
Institute for Research in Biomedicine (IRB Barcelona), Barcelona Institute of Science and Technology (BIST), 08028 Barcelona, Spain
2
Catalan Institution for Research and Advanced Studies (ICREA), 08010 Barcelona, Spain
3
Cambridge Institute of Science, Altos Labs, Granta Park, Cambridge CB21 6GP, UK
*
Author to whom correspondence should be addressed.
Pharmaceutics 2023, 15(2), 352; https://doi.org/10.3390/pharmaceutics15020352
Submission received: 21 November 2022 / Revised: 22 December 2022 / Accepted: 26 December 2022 / Published: 20 January 2023
(This article belongs to the Special Issue Advances in Mitochondria-Targeted Drug Delivery)

Abstract

:
Ageing is accompanied by a progressive impairment of cellular function and a systemic deterioration of tissues and organs, resulting in increased vulnerability to multiple diseases. Here, we review the interplay between two hallmarks of ageing, namely, mitochondrial dysfunction and cellular senescence. The targeting of specific mitochondrial features in senescent cells has the potential of delaying or even reverting the ageing process. A deeper and more comprehensive understanding of mitochondrial biology in senescent cells is necessary to effectively face this challenge. Here, we discuss the main alterations in mitochondrial functions and structure in both ageing and cellular senescence, highlighting the differences and similarities between the two processes. Moreover, we describe the treatments available to target these pathways and speculate on possible future directions of anti-ageing and anti-senescence therapies targeting mitochondria.

1. Introduction

Cellular senescence is recognized as a hallmark of ageing [1] and cancer [2]. Cellular senescence is a response to severe damage or stress characterised by the inability to proliferate and by a robust secretion of high amounts of inflammatory, fibrogenic and mitogenic factors, collectively known as the senescence-associated secretory phenotype (SASP) [3,4]. Cellular senescence serves important physiological functions, most notably to initiate tissue repair and to prevent the outgrowth of potentially oncogenic cells. After exerting their beneficial functions, senescent cells are normally cleared by the immune system. However, with ageing, senescent cells are not efficiently cleared and accumulate contributing to many diseases. Indeed, the accumulation of senescent cells can lead to disruption of tissue functionality, and limit the regenerative potential of adult stem cells by damaging the stem cell niches within the affected tissue [5]. Although still speculative, the SASP, rather than the physical presence of senescent cells, could be the main pathological agent of cellular senescence.
For the purposes of this review, we will use the terms “cellular senescence” and “cellular ageing” to describe two different cellular processes. We will use “cellular ageing” when referring to what happens to cells when the organism ages, and we will use “cellular senescence” when referring to what happens to cells when they reach a threshold of severe damage and undergo adaptations that profoundly change cellular biology. “Cellular ageing” involves progressive changes that deteriorate cell functions, often over the span of years, but without drastically changing their main biological properties. In contrast, “cellular senescence” involves a dramatic transformation of the cell biology in a process that typically takes about 7–10 days (summarised in Figure 1). While “aged cells” are suboptimal versions of their younger counterparts, “senescent cells” are very different versions with multiple aspects of their biology altered, including major chromatin reconfiguration, remarkable expansion of the lysosomal compartment, and increased autophagy [4]. In the following sections, we will try to summarise and compare how mitochondrial biology changes during “cellular ageing” and upon “cellular senescence”.
There is a growing interest in therapeutically targeting both “aged cells” and “senescent cells”. The goals are different: in the case of “cellular ageing”, it would be ideal to find interventions that (1) slow down the pace of deterioration of cells, or (2) reverse the accumulated damage. In the case of “cellular senescence”, the goals are (1) to eliminate senescent cells either by driving them into apoptosis (senolytic strategies) or by stimulating their immuneclearance, or (2) to reduce the production of their main pathological mediator, namely, the SASP. It is important to emphasise that the senescence-targeting therapies do not prevent or interfere with the de novo implementation of senescence (which is an important anti-oncogenic barrier), but act on pre-existing senescent cells that have not been naturally cleared by the immune system. In this review, we will discuss strategies particularly targeted to the mitochondrial biology of “cellular ageing” and “cellular senescence”.
Mitochondria are ubiquitous intracellular organelles essential for multiple cellular functions. Indeed, these organelles are central in the metabolic processes involved in ATP and energy production but are also implicated in calcium homeostasis, ROS signalling, apoptosis, haem and iron-sulphur clusters synthesis, inflammation, and epigenetics regulation. Numerous studies have shown how mitochondria in aged and senescent cells undergo heavy structural changes and general functional decline (summarised in Figure 2), suggesting that this organelle might be a favourable target to tackle ageing-associated diseases. However, the understanding of these modifications is still limited and additional and improved experimental and clinical data is required before applying this knowledge to the healthcare practice.

2. Mitochondria as Inflammation Triggers

Ageing is associated with increased inflammation and activation of the innate immune system. This condition is known as “inflamm-ageing” and is characterised by chronic activation of JAK-STAT signalling in the circulating immune cells of elderly patients [6], activation of the NLRP3 inflammasome [7,8,9], and higher circulating levels of inflammatory mediators such as C-reactive protein, IL-6, and fibrinogen [10]. A leading hypothesis for the origin of “inflamm-ageing” is the build-up of senescent cells with ageing. An important support for this hypothesis comes from experiments in which aged mouse blood is transferred to young animals, which results in features of accelerated ageing. Interestingly, previous treatment of the old donors with senolytic agents reduced “inflamm-ageing” after blood exchange, and the old blood lost its pro-ageing activity [11]. In humans, senolytic treatments also reduce the “inflamm-ageing” of patients suffering from lung fibrosis [12] or chronic kidney disease [13].
Importantly, mitochondria of senescent cells are known to play a key role in triggering the SASP. In particular, depriving senescent cells of mitochondrial DNA [14] or mitochondria altogether [15] seriously compromises the SASP. The detailed mechanisms connecting the mitochondria of senescent cells with the SASP are still unknown. We speculate that they could be similar to the mechanisms connecting dysfunctional mitochondria with inflammation [16]. These may include the release of cytosolic and/or extracellular mitochondrial DNA (mtDNA), mitochondrial double-stranded RNA, N-formyl peptides (a sub-product of mitochondrial protein translation), and phospholipid species such as cardiolipin, enriched in the inner mitochondrial membrane (IMM) [17,18,19]. The most studied of these components is mtDNA which will be analysed in depth in the next section. Apart from mtDNA, it is worth mentioning that formyl peptides can be released extracellularly by damaged mitochondria and activate neutrophils by engaging their formyl peptide receptor-1 (FPR1) [18]. Specific FPR1 antagonists have been generated and used to treat brain tumours and neurodegeneration [20,21], suggesting a potential use to fight age-driven inflammation.

3. Mitochondrial DNA

As a reflection of their bacterial origin, mitochondria contain their own genetic material, mtDNA. This circular double-stranded molecule only counts 16,569 base pairs [22], but it is present in hundreds of copies in each cell and its contribution to organismal ageing has been extensively discussed. The initial hypothesis was that the accumulation of mtDNA mutations with age might directly contribute to the decline of mitochondrial functions. Indeed, compared to nuclear DNA, mtDNA is in close contact with the electron transport chain, the principal producer of reactive oxygen species (ROS) that can induce oxidative DNA damage, is less protected, and its repair mechanisms are far less sophisticated [23]. Supporting this hypothesis, mtDNA deletions and mutations are detected in tissues from aged animals and humans [24,25,26,27]. Moreover, the generation of an mtDNA mutator mouse model, which expresses a proof-reading-deficient version of the mitochondrial DNA polymerase G and accumulates mtDNA mutations at vastly increased rates, showed reduced lifespan and premature onset of ageing when the mutation is in homozygosity [27,28]. However, heterozygous DNA polymerase G mutant mice show normal ageing despite huge levels of mtDNA mutations and the amount of mtDNA mutations that accumulate during natural ageing is far lower than in these mutant mice [29]. As is true for oncogenic nuclear DNA mutations, it has been shown that mtDNA mutations accumulate in several human tumours, particularly in genes encoding for complex I subunits. These mutations favour oncogenesis by inducing metabolic remodelling, accelerating cell proliferation, and reducing apoptosis at least in certain tissues [30,31]. In this case, cellular senescence could be initiated as a defence mechanism to suppress the development of a tumour [23].
Less controversial, instead, is the role of mtDNA on organismal ageing when released outside of the mitochondrial matrix, both into the cytosol or extracellularly. This process increases in senescent cells [32] and is now a hot topic of research because of its impact on inflammation and immune responses due to the mitochondrial genome’s bacterial-like nature. Cytosolic escape of mtDNA triggers the cGAS-STING-NLRP3 axis, a key process in response to cellular stress [33,34], and consequently activates the interferon regulatory factor 3 (IRF3) or the transcription factor family nuclear factor kB (NF-κB) pathway, major players in inflammation and antiviral response [35,36,37]. The mechanisms involved in mtDNA escape are still debated, which makes finding direct inhibitors complicated. An indirect strategy has been to reduce mtDNA release as a downstream effect of other treatments including reducing oxidative stress with melatonin supplementations [38], or reducing the opening of the mitochondria permeability transition pore (mPTP), a possible exit way for mtDNA outside of the matrix, with cyclosporin A [39]. These studies, however, have not yet offered a clear mechanistic explanation of how these compounds inhibit mtDNA release. Another approach is to act directly on the cGAS pathway. A variety of small molecules able to inhibit cGAS activation have been successfully identified and developed in treating autoimmune diseases, such as RU.521, which competitively binds to cGAS catalytic pocket with cGAS substrates ATP/GTP [40], Cu-32 and Cu-76, that prevent cGAS dimerisation and subsequence activation [41], and additional small molecules identified through screenings: G140/G150 [42] and PF-06928215 [43]. When mtDNA is released in the extracellular environment and reaches the bloodstream, instead, it can be taken up by immune cells, such as neutrophils and macrophages, by endocytosis, and activate the Toll-like receptor 9 (TLR9), a pillar of antibacterial and antiviral responses [44,45]. The downstream effect is the activation of NF-κB, the secretion of tumour necrosis factor-ɑ (TNF-ɑ), and the expression of the pro-interleukin-1β and -18 in both tissue-resident macrophages and circulating leukocytes [46,47], leading to the recruitment of other immune factors. Circulating mtDNA appears to increase gradually with age after the fifth decade of life and to be strictly associated with inflammatory status [48]. Indeed, levels of circulating mtDNA correlate with serum inflammatory markers [48]. For this reason, designing new therapeutic strategies against circulating mtDNA, or the receptors it binds, could be relevant in the future treatment of ageing-associated diseases.

4. Mitochondrial Life: Biogenesis, Dynamics and Mitophagy

Mitochondria are very dynamic organelles. They undergo constant fusion and fission events to create a specific network able to accommodate the cellular energy demands and metabolic state, allow transport, and favour the selective removal of damaged mitochondria through a process known as “mitophagy” [49]. While elongated mitochondria permit the sharing of metabolites, proteins, and mtDNA and enhance cell survival, mitochondrial fragmentation is often associated with motility or mitophagy, and, in more extreme cases, is a sign of mitochondrial dysfunction and cell death [50].
Mitochondrial dynamics change with organismal ageing and cellular senescence. Ageing in flies and mammalians is characterised by enlarged mitochondria, irregular cristae shape and size, and a decrease in mitochondrial number [51,52]. In addition, senescent cells, are characterised by a very elongated and branched mitochondrial network [53,54]. A possible cause for this fusion-oriented phenotype is the reduced expression of FIS1 and DRP1, two of the proteins involved in promoting mitochondrial fission, during senescence [55]. This downregulation might reflect an attempt to dilute and re-arrange matrix content between healthy and damaged organelles or to resist apoptosis [56]. Indeed, there is a tight interconnection between apoptosis and mitochondrial dynamics since DRP1 relocates from the cytosol to mitochondria during cell death, resulting in mitochondrial fragmentation, loss of membrane potential, and cytochrome c release [57,58]. When this protein’s activity is inhibited, the development of a senescent phenotype is favoured [59], while the induction of Drp1p expression in Drosophila midlife prolongs both life- and health-span via improved mitochondrial respiration and autophagy [60]. Similarly, the induction of mitochondrial fission in the intestine of both C. elegans and flies increases longevity [61]. Upregulation of mitochondrial fission could therefore ameliorate senescence-related phenotypes.
Another common hallmark of organismal ageing and cellular senescence is impaired mitophagy [62,63], which leads to the accumulation of dysfunctional organelles, as observed in both old rats and humans cells [64,65], and also in senescent cells in vitro and in vivo [66,67]. Mitophagy decline might result from several molecular mechanisms. Firstly, defects in lysosomal function or lysosomal overload might prevent lysosomal enzymes from targeting autophagosomes, leading to defective removal of dysfunctional mitochondria [61,68]. In aged cells from old tissues as well as in senescent cells, lysosomes show reduced activity and accumulation of undegraded material [69]. Secondly, an overall defect in cellular autophagic capacities could explain the deficiencies in mitochondrial clearance. The mTOR-autophagy axis is affected during ageing, and senescent cells show elevated mTORC1 activity [15,66,70], becoming unresponsive to starvation signals [71]. This hypothesis is supported by studies where autophagic flux was restored and age-related conditions were prevented in rodents, dogs, non-human primates, and humans after treatment with two mTOR inhibitors: rapamycin [72,73,74] and AZD80055 [75]. In addition, impaired mitophagy might derive from the mitochondrial dynamics defects previously described. Indeed, the efficiency of this process relies on the ability of the organelle to undergo fission and segregate the segment of the network that needs to be eliminated [76]. Moreover, both fission and mitophagy efficiency can be worsened with ageing because of a reduced expression of PINK1 [77], which, together with Parkin, is the main actor of the mitochondrial clearance pathway [78].
While mitochondrial turnover decreases in both aged and senescent cells, mitochondrial biogenesis slows down during ageing [79] but appears to increase during senescence, leading to a rise in respiration and ROS production [80]. The underlying reasons still have to be completely elucidated, but a possible explanation for this difference is the expression level of a key regulator of biogenesis, PGC-1α, which is higher in senescent cells [80] and reduced in aged animals [81]. However, despite different screenings to identify compounds able to modulate PGC-1α transcription, results are still elusive [82,83] and alternative approaches are needed. Another proposed way to pharmacologically modulate mitochondrial biogenesis is targeting the AMP-activated protein kinase (AMPK) pathway. In aged animals, chronic AMPK inactivation is associated with a marked decrease in mitochondrial biogenesis [84], while reduced AMPK activity correlates with ageing-related insulin resistance and insufficient intracellular fat oxidation [85]. On the other hand, when aak-2, the worm equivalent of AMPK, is overexpressed in C. elegans, it results in an increased lifespan [86]. Thus, chronic activation of AMPK via metformin treatment, a compound already used in the clinic for the treatment of type 2 diabetes, has been proposed as a strategy for slowing ageing [87]. However, while preliminary studies showed promising results, it is unclear if the benefits come from its direct action on AMPK or indirect effects on cellular metabolism and glycaemia, reduction of oxidative stress, or protective effects on the endothelium and vascular function [88]. The molecular mechanism of metformin is also debated, since metformin can activate AMPK as a downstream effect of its interaction with various proteins and pathways, including mitochondrial complex I, the nuclear receptor NR4A1, and the endosomal/lysosomal v-ATPase [89,90]. Moreover, the effects of metformin on mice longevity are not robust across different laboratories. While certain studies showed increased lifespan [91], others reported no changes [92] or even decreased longevity [93], highlighting the importance of comprehensive testing before the introduction of metformin-based anti-ageing therapies.
Interestingly, the role of AMPK activity and its function in the regulation of mitochondrial biogenesis in cellular senescence is less clear. A study in immortalised human fibroblasts expressing SV40 large T antigen showed that inactivation of large T resulted in an increase in AMPK activity that directly contributed to the implementation of senescence [94]. In contrast, a study in H2O2-induced senescent murine fibroblasts found AMPK inactivated [95]. Since AMPK is central in multiple cellular pathways, additional research could clarify the observed differences between models, and unravel additional molecular mechanisms involved in the establishment and/or maintenance of senescence.

5. Mitochondrial Unfolded Protein Response

Mild mitochondrial stress can be beneficial for longevity [96,97] and a reason for this hormetic effect could be the activation of the mitochondrial unfolded protein response (UPRmt) [98]. The UPRmt, similarly to the endoplasmic reticulum unfolded protein response (UPRER) and the cytoplasmic heat shock response (HSR), is capable of initiating a broad-range transcriptional response that not only is involved in the refolding of mitochondrial matrix proteins, but also in ROS defences, metabolic changes, regulation of iron-sulphur cluster assembly, and modulation of the innate immune response [99,100,101]. Lead UPR factors include the heat shock protein 60 and 10 (HSP60 and HSP10), mitochondrial heat shock protein 70 (mtHSP70), Lon peptidase 1 (LONP1) and caseinolytic protease (ClpP).
While it has not been clearly described if, when, and how a decline in UPRmt functionality takes place during the ageing of the organism, it is largely accepted that UPRmt activation has a beneficial effect on longevity since it promotes cell survival and the recovery of the mitochondrial network and cellular function. However, the beneficial activation of UPRmt might not be a viable route in tackling ageing therapeutically. Indeed, UPRmt triggering after exposure to mitochondrial stress shows strong responses only during development [102,103], while it appears less active in later stages of life when there is no reported increased lifespan as a response to stressors [96,104,105]. This timing limitation has been justified by the fact that mitochondrial perturbations early in life induce widespread changes in chromatin structure through histone H3K9 di-methylation and long-lasting effects on gene expression [106]. The transcription of target UPRmt genes is subject to epigenetic regulation by histone3-specific methylation and is therefore influenced by those stresses that occurred during development [106,107] while being less sensitive to treatments in aged organisms.
Similarly, limited experimental data are available about the UPRmt and senescence. In senescent hepatocytes, most of the UPRmt factors levels were found significantly reduced and the pathway compromised [108], suggesting that UPRmt targeting could have a role in preventing senescence. Similarly, experiments in senescent human lung fibroblasts showed a reduced ability to cope with the accumulation of mitochondrial unfolded proteins [109]. Additional research is needed to investigate this hypothesis.

6. Metabolism and Electron Transport Chain

Mitochondria’s primary function is to be the “powerhouse of the cell”. Thanks to the activity of the electron transport chain (ETC, composed of four enzymatic complexes embedded in the IMM), ATP is formed from adenosine diphosphate and inorganic phosphate and becomes available for the cell to be used as “energy currency” [110]. Analyses of mitochondrial function in skeletal muscle samples from older subjects showed a strong decline in mitochondrial respiratory capacity and a reduction in ATP amount [111,112]. At the same time, ROS produced by the ETC enzymes increase in aged animals [24]. Similarly, while experiments on senescent fibroblasts showed an increase in mitochondrial mass and abundance of tricarboxylic acid (TCA) cycle metabolites [113,114], the efficiency of the ETC and ATP production appeared reduced, leading to decreased mitochondrial membrane potential, increased proton leak and generation of ROS [115,116]. Because of this defect, senescent cells appear to undergo a metabolic switch, increasing dependence on glycolysis [117], and fatty acid oxidation [118]. As discussed previously, the cause of the ETC damage during both ageing and the establishment of cellular senescence is probably the combination of increased ETC machinery malfunction due to progressive damage and a decline in the removal of dysfunctional mitochondria.
For all these reasons, a direct or indirect amelioration of the energetic capacity and functionality of mitochondria in ageing and senescent cells could improve patients’ symptomatology. This result could be achieved by reducing stressful or damaging conditions, such as ROS and calcium accumulation, impaired mitophagy or altered dynamics. However, the ETC itself can be directly targeted. For example, both aged tissue and senescent cell exhibit low levels of Coenzyme Q10 (CoQ10), the ETC carrier that collects electrons from complexes I and II and delivers them to complex III [119,120]. This deficiency can lead to electron leakage, loss of membrane potential, ROS production, and reduced ETC efficiency [121]. In mice with accelerated ageing, CoQ10 supplementation improves complexes I and IV activity and OXPHOS efficiency and decreases ROS generation, slowing down the progression of ageing-related symptoms and preventing ageing [122]. CoQ10 supplementation could have beneficial effects also against the development of cellular senescence, as demonstrated in mesenchymal stem cells [123] and H2O2-induced senescent HUVECs [124].
Furthermore, favouring alternative energy-producing pathways such as β-oxidation can show beneficial effects on lifespan and metabolism [125]. Direct administration of fatty acids, including α-linolenic or omega-3 fatty acid, indeed, promotes higher mitochondrial energy production, mitochondrial biogenesis, and oxidative stress reduction [126]. This hypothesis was tested also in the clinic, where humans aged 65 to 85 showed increased mitochondrial protein synthesis and significantly reduced mitochondrial oxidative stress after being treated with ω-3 polyunsaturated fatty acids [127]. Another proposed treatment able to increase β-oxidation is the supplementation of 17α-estradiol, a weak endogenous steroidal oestrogen. 17α-estradiol improved metabolic parameters and slowed ageing in male mice, but did not show significant effects in females [128]. However, conclusive clinical trials about the impact of these treatments on human health and longevity are still missing.

7. NAD+ Levels

Nicotinamide adenine dinucleotide (NAD+) is one of the most common metabolites in the human body and an indispensable cofactor involved in several redox reactions. Most NAD+ functions as a redox carrier, receiving electrons from metabolic processes such as glycolysis, Krebs cycle and β-oxidation, and forming NADH. NADH is then used to transfer electrons to complex I in the ETC [129]. Approximately 10% of cellular NAD+, instead, can be phosphorylated by NAD+ kinases into NADP+ and used for anabolic reactions, such as lipid and nucleic acid syntheses, which require NADPH as an electron donor [130,131].
NAD+ levels decline with age in various tissues, and this reduction correlates with the development of ageing-associated diseases such as muscle loss and diabetes [132,133,134]. A drop in NAD+ levels, indeed, associates with mitochondrial dysfunction in both calorie-rich diets and ageing, whereas NAD+ repletion with precursors such as nicotinamide riboside and nicotinamide mononucleotide can reverse this process, improving mitochondrial respiration and increasing ETC subunits expression [132,133,135]. Nonetheless, decreased NAD+/NADH ratios or total NAD+ levels can drive senescence and cell cycle arrest, but also influence the SASP phenotype [14,136]. For example, cells that underwent mitochondrial dysfunction-associated senescence are characterised by lower NAD+/NADH ratios, AMPK-mediated p53 activation, and reduced IL-1-associated SASP [14]. The complexity of the role of NAD+ in the establishment of senescence is linked to its activity as a cofactor of two important protein families: poly-ADP-ribose polymerase (PARP) and sirtuin family proteins (SIRTs). PARP has a double and contradictory role in the establishment of senescence, making it difficult to use this protein as an anti-ageing target. On one side, indeed, PARP inhibition or depletion leads to single-strand breaks, cell cycle arrest, and cellular senescence [137,138,139]. On the other side, its activity promotes NF-kB activation and secretory phenotypes in senescent cells [140]. Consistently, a decrease in NAD+ amount by inhibition of Nicotinamide phosphoribosyltransferase, responsible for the NAD salvage pathway, promotes both cell cycle arrest and suppresses SASP [14,136]. While all the mentioned studies have been conducted on human cells, the use of different cell models (keratinocytes, ovarian cancer, colorectal cancer, melanoma and breast cancer cells) and the use of different senescence triggers might explain these disparities. Likewise, SIRTs play a role in senescence and SASP generation. The mitochondrial sirtuin SIRT3 is critical in the elimination of intracellular ROS and the maintenance of oxygen metabolism balance [141]. Interestingly, the increased oxidative stress that results from its depletion can induce different senescence markers but suppresses SASP secretion. Finally, it has been recently discovered that NAD+ levels can be restored in stressed or damaged cells by a cytosolic complex of enzymes that transfers electrons from NADH to NADP+, and this reaction can prevent cellular senescence [142].
For all these reasons, the modulation of NAD+ levels and the downstream molecular pathways have been largely studied as a potential target for anti-ageing therapies. The most direct proposed intervention was to increase NAD+ levels by dietary, via supplementation of NAD+ precursors, such as nicotinic acid, nicotinamide riboside, nicotinamide mononucleotide, and tryptophan, or improving NAD+ bioavailability through exercise and caloric restriction. However, while these therapies have already shown promising results in clinical trials, it is necessary to remember that NAD+-depleting drugs have an anti-tumoural effect, and the long-term boosting of NAD+ might increase the risk of developing cancer. Consistent with this observation, nicotinamide mononucleotide treatment accelerates pancreatic cancer progression inducing an inflammatory environment [136]. Another strategy could be the reduction of PARP1 and CD38 activity, which consume NAD+ [143,144]. US-FDA-approved PARP inhibitors, such as niraparib, rucaparib and olaparib, are already available and used to treat cancers, including prostate, breast and ovarian, through disrupting DNA repair and replication pathways [145,146,147]. Similarly, CD38 inhibitors have been proposed, such as apigenin, quercetin, luteolin, kuromanin, luteolinidin, and 78c [148]. 78c, for example, is a highly specific CD38 inhibitor which showed promising results in reversing NAD+ decline during ageing and improving age-associated cardiac and muscle function and glucose tolerance [143]. The opposite tactic, instead, could be stabilising NAD+ levels by increasing NAD+-biosynthesis enzymes activity [149] or preventing the escape of intermediates [150].
Finally, SIRTs themselves could represent a powerful tool in anti-ageing therapies. Overexpression of SIRT3 showed beneficial effects on ageing and senescence hallmarks [151] and significantly activates mitophagy [152]. Increased SIRT3 activity can be reached through calorie restriction [153] or compounds such as adjudin, a derivative of lonidamine [154], which was described attenuating cellular senescence markers in hydroxyurea-treated MEFs [155]. However, the potential of this pharmacological approach remains to be validated in clinical conditions.

8. Matrix Calcium

Calcium (Ca+2) is an ion that participates in a wide variety of cellular functions, being an intracellular regulator of many physiological processes. Mitochondria and Ca+2 are strictly interconnected. On one hand, the cell benefits from the mitochondrial ability to buffer Ca2+, shaping the cytosolic Ca2+ signal [156,157]. On the other, Ca2+ is fundamental for normal mitochondrial functions since it activates pyruvate, isocitrate, and 2-oxoglutarate dehydrogenases, involved in the TCA cycle, stimulating mitochondrial respiration and ATP production [158]. Excessive and prolonged accumulation of mitochondrial Ca2+, often in combination with increased production of ROS [159], however, can be toxic and regulate cell death through the stable opening of the mPTP [160]. This phenomenon leads to mitochondrial swelling, metabolism impairment, alterations in the matrix content, membrane potential drop, and apoptosis [161,162,163]. Because of this, calcium movements within organelles are dynamic but strictly controlled. Its route starts in the endoplasmic reticulum (ER), which releases calcium through ryanodine receptors and inositol 1,4,5-triphosphate receptors (Ins(1,4,5)P3Rs). In particular, Ins(1,4,5)P3Rs are enriched at the ER-mitochondria contact sites, areas of proximity but not fusion, between the membranes of the two organelles [164], and are activated by IP3 binding. On the receiving side, mitochondria express VDAC1 in the outer mitochondrial membrane (OMM) and the mitochondrial calcium uniporter complex (MCUC) in the IMM that uptakes Ca2+ into the matrix.
Recent observations pointed out that a rise in intracellular Ca2+ contents could be a new hallmark of ageing and cellular senescence, both at a cytosolic level and in the intracellular organelles [165,166], leading to chronic mitochondrial stress and cellular toxicity. Evidence shows increased activation of Ins(1,4,5)P3Rs in the ER membrane of senescent cells and consequent amplified uptake of Ca2+ through MCU channels, possibly due to a decreased expression of the transient receptor potential cation channel subfamily C member 3 (TRPC3) [167], a controller of mitochondrial Ca2+ load. Re-expression of TRPC3, indeed, diminished mitochondrial Ca2+ load of these cells and promoted escape from oncogene-induced senescence. Nonetheless, additional alterations in the mitochondrial calcium machinery structure or activity cannot be excluded and should be further investigated.
Mitochondrial calcium concentrations could be, therefore, an attractive target in both anti-ageing and anti-senescence therapies. While calcium modulation in senescent cells is still poorly studied, it has been proposed to use specific MCU inhibitors, such as Ru360 or Ru265, to reduce the entry of Ca2+ in the matrix of ageing cells, and consequentially mitochondrial stress. These compounds have been largely characterised in vitro and in vivo, and have shown promising protective activity in reperfusion [168] and hypoxic/ischemic brain injury [169] animal models. Recent work in C. elegans started testing this hypothesis in the context of ageing and reported that pharmacological or genetic inhibition of MCU was sufficient to improve muscle ageing and dystrophy, corroborating this hypothesis [170]. In addition, future studies should investigate other possible therapeutic targets, including the modulation of ER-mitochondria tethering and Ins(1,4,5)P3Rs activity, the role of other subunits of the MCU complex, and other mechanisms of mitochondrial calcium influx and efflux in ageing and senescence.

9. Reactive Oxygen Species

Mitochondria are a major source of ROS, which are primarily the result of the inefficient transfer of electrons through the ETC. According to the mitochondrial free radical theory of ageing, ROS are both a central cause and a consequence of ageing. Indeed, reportedly they increase with age because of a decline in ETC capacity, respiratory complexes dysfunctions, and a decrease in ROS scavenging enzymes, whereas, on the other side, ROS accumulation increases levels of oxidized lipids and proteins, induces mtDNA mutations, and further deteriorates the ETC [171,172,173]. In cellular senescence, a similar pattern can be observed: ROS contribute to cellular senescence onset inducing oxidative damages [174] and inhibiting autophagy [175], while the chronic accumulation of these species establishes a vicious cycle of mitochondrial and cellular stress.
Thereby, it would be easy to believe that the elimination or drastic reduction of ROS would be an effective anti-age therapy. While this reasoning is not intrinsically wrong, the situation is more complicated. Mitochondria-targeted antioxidant drugs such as plastoquinone derivatives [176,177] or MitoTEMPO [178,179], and endogenous indoleamine melatonin [180], validate this theory, showing increased lifespans in mice and flies and generally improved mitochondrial functions. Likewise, genetically modified mice with reduced ROS production show delayed ageing [181,182]. On the contrary, however, additional reports raised doubts about the free radical theory of ageing and the use of antioxidant therapies. In mice, the overexpression of major antioxidant enzymes such as copper-zinc superoxide dismutase (CuZnSOD or SOD1), catalase, or manganese superoxide dismutase (MnSOD) did not increase longevity [183], while deletion of mitochondrial matrix SOD increased mtDNA damage and cancer incidence but did not accelerate ageing [184,185]. In C. elegans, instead, loss of superoxide dismutase enzymes could even extend lifespan [186]. Finally, the use of certain antioxidants on proliferating cells can favour cellular senescence, by inducing proliferation arrest, DNA damage and chromosomal abnormalities [187].
These controversies highlight that cellular ROS are not only a damaging “waste product” of mitochondrial activity but have critical functions in cellular life. Emerging evidence, indeed, show the importance of ROS in cellular signalling. For example, H2O2 generated from superoxide produced by mitochondria and NADPH oxidases [188,189] mediates the oxidation of cysteine residues [190], causing allosteric changes within important signalling proteins and modifying their behaviour. Also, H2O2- can promote tyrosine phosphorylation by activating protein tyrosine kinases. Other evidence suggests that ROS signalling is required for the maintenance of tissues since it can activate cellular stress pathways to diminish tissue degeneration and promote healthy ageing [191]. Besides, ROS are also essential for stem cell differentiation, as observed in different in vivo models: murine hematopoietic stem cells in a mouse model with reduced ROS levels because lacking AKT1 and AKT2 showed compromised differentiation [192], while in Drosophila hematopoietic progenitors, increasing ROS triggers differentiation while decreasing ROS impairs it [193]. Experiments in humans also reported the importance of mitochondrial ROS in muscle differentiation and the differentiation into adipocytes of bone marrow mesenchymal stem cells [194,195]. Accordingly, reduced ROS levels decrease the regenerative capacity of neural stem cells and spermatogonial stem cells [196,197]. However, a few conflicting studies claimed that a rise in ROS might harm stem cell function [198,199], suggesting that there might be exceptions to the rule.
The ROS situation, therefore, is not black or white, but the type of ROS, their localisation, and their concentration collectively determine whether redox signalling or oxidative stress-induced damage occurs. Therefore, the problem in ageing and senescence is not ROS on its own, but a dysregulated and atypically high production of them, paired with redox dyshomeostasis. In conclusion, while aberrant ROS generation likely plays a role in age-related pathologies, antioxidant therapies will need to be carefully modulated to be both effective and not toxic.

10. Mitochondrial Permeability Transition Pore

The mPTP is a transmembrane protein or complex that controls mitochondrial permeability and can induce cell death after various stresses, including oxidative stress, adenine nucleotide depletion, increased phosphate concentration, and high mitochondrial calcium [200]. The opening of the pore leads to mitochondrial swelling, uncontrolled diffusion of molecules under 1500 Da across the IMM, and sustained loss of mitochondrial membrane potential [162]. The permeability of the mitochondrial membranes is central in the decision between cell survival or death, and what type of cell death, since the activation of the pore has been associated with both apoptosis and necrosis [201]. While many potential structural components of the mPTP have been proposed, including ATP synthase, adenine nucleotide translocase (ANT), the outer membrane voltage-dependent anion channel (VDAC), and the phosphate carrier (PiC), the exact subunit composition of the pore is still debated [201]. Its regulatory pathways are also unclear. Cyclophilin D (CypD) is the only clearly described mPTP regulator: it controls mPTP opening by sensitising it to calcium, inorganic phosphate, and ROS, while other stimuli can activate mPTP opening via CypD-independent pathways [200,202].
A common feature of old tissues and senescent cells is an increased concentration of mitochondrial Ca2+ and ROS, two conditions that can directly stimulate the opening of the mPTP [203,204], and decreased levels of NAD+ [132,133,134], which results in low levels of SIRT3 activity and thereby high levels of the active, acetylated, form of CypD [205]. Together, these alterations (high Ca2+, high ROS and low NAD+) make the mitochondria of old and senescent cells highly prone to mPTP opening. Studies in muscles from aged humans and rats reported reduced mitochondrial calcium retention capacity and sensitisation of the mPTP opening, leading to apoptosis [64,206]. Similarly, enhanced susceptibility to mPTP opening during ageing was found in the brain [207,208], the liver [208,209], and lymphocytes [210]. The consequences of increased mPTP are the collapse of mitochondrial membrane potential, reduced mitochondrial respiratory function, the release of mitochondrial Ca2+ and cytochrome c, and enhanced ROS generation [202,211], all events that have been linked to ageing. On the contrary, our knowledge about mPTP opening in senescent cells remains speculative and its potential use as a therapeutic target needs to be investigated.
In different degenerative diseases, typically associated with ageing, the inhibition of the mPTP has been tested as a protective strategy to preserve cell survival, showing encouraging results. In mouse models of Parkinson’s disease and amyloid lateral sclerosis, for example, the prevention of the pore opening by CypD depletion or pharmacological inhibitors of the mPTP showed delayed onset of disease and extended lifespan [212,213,214]. The same happened in ageing-related bone loss, where CypD knock-out mice showed enhanced resistance to osteoporosis [215].
In most diseases associated with cellular senescence, instead, the goal is the opposite, namely, to eliminate senescent cells. While an increase in mPTP activation could lead to extensive toxicity and damage also proliferative cells, a deeper understanding of the mPTP activity and regulation could offer new opportunities for intervention. Indeed, while pore activation has traditionally been considered a death sentence for the cell, additional evidence proved that the mPTP can open in two different ways: permanently or transiently. While the sustained opening leads to cell death, the temporary activation or “flicker” of the pore [216] can have beneficial effects or induce protective pathways. It can allow calcium, ROS or other molecules release or exchange between the mitochondrial matrix and the cytosol, activate rescue pathways, or act as a signalling event. Since Ca2+ and ROS levels are elevated in aged and senescent cells, a transient variation of the mitochondrial permeability could represent a weapon against overload and toxicity, while its inhibition could target a specific vulnerability of these cells and have a senolytic effect. Transient opening and moderate loss of mitochondrial membrane potential could also correlate with activation of the UPRmt, which contributes to health and longevity, as previously discussed [102,107], favouring cell survival. The mPTP has, therefore, two highly different functions, one that rescues the organelle and the cell in cases of moderate stress and one that condemns it when its condition is beyond repair. Future research on mPTP targeting strategies should keep this important difference in mind and, ideally, identify ways to discriminate between the two pathways.

11. Apoptosis

Apoptosis is the process that leads to a controlled and programmed cell death, which can occur as a response to various damages or stresses, such as DNA damage, oxidative stress, immune reactions and absence of certain growth factors, hormones and cytokines, or as a natural part of embryonic development and ageing [217]. The apoptotic program can be initiated by different triggers and follow different signalling pathways, which generally share the activation of initiator caspases (as caspase 8 and 9) and culminate in the activation of executioner caspases (as caspase 3 or 7) to finally induce DNA fragmentation, degradation of cytoskeletal and nuclear proteins, cross-linking of proteins, and formation of apoptotic bodies. Mitochondria are responsible for the initiation of a key apoptotic pathway, as a result of internal stress or damage signals, which consists of the mitochondrial outer membrane permeabilisation (MOMP), regulated by the Bcl-2 family of proteins. The Bcl-2 family includes both pro- and anti-apoptotic proteins, which balance their activity in cells as needed. Upon elevated stresses, the pro-apoptotic proteins BAX and BAK oligomerise at the OMM [218,219], where they induce the release of cytochrome c and other proteins from the intermembrane space into the cytosol [220]. Once in the cytosol, cytochrome c induces the downstream activation of apoptotic protease activating factor-1 (Apaf-1), caspase 9, and finally, caspase 3, starting the execution pathway [221,222]. To avoid the undue activation of this suicide pathway, another group of proteins, including Bcl-xL, Bcl-2, and Bcl-W, prevent the oligomerisation of BAX/BAK [223,224].
The positive influence of the anti-apoptotic side of this family on healthy ageing was confirmed by the fact that Bcl-xL was found overexpressed in centenarian subjects [223]. Moreover, in “old” fibroblasts with high passage number, the level of anti-apoptotic proteins increases dramatically after UV stress and favours the development of a senescent phenotype, while in UV-damaged “young” cells with low passage number, the pro-apoptotic pathway is preferred [225]. These results suggest that activation or upregulation of the Bcl-xL pathway could be a valid anti-ageing strategy, but it can also help preserve damaged cells in older tissue, which are more prone to develop a senescent phenotype. Since senescent cells’ survival depends more than their proliferative counterparts on the anti-apoptotic activity of these members of the Bcl-2 family, their downregulation or inhibition has been exploited for their senolytic effect. One of the few and most used senolytic agents available to date is, indeed, Navitoclax (ABT-263), which induces apoptosis of senescent cells by inhibition of Bcl-W and Bcl-xL [226,227]. Preliminary data from clinical trials using Navitoclax in combination with Ruxolitinib in patients with myelofibrosis [228,229] showed encouraging outcomes, but further studies are necessary to fully evaluate the potential of this novel combination. Finding additional senolytic targets or strategies to make the already existing ones more specific is, therefore, one of the most active areas in ageing research.

12. Epigenetic Regulation

Although the DNA code in our cells carries the genetic information, the epigenome is responsible for the accessibility, stability and regulation of that valuable information, connecting the genotype with the phenotype [230,231]. Epigenetic changes are reversible, can be driven by external or internal influences, and represent a key mechanism behind cellular alterations during ageing. In mammals, older individuals tend to present with CpG hypomethylation, especially at repetitive DNA sequences [232,233,234], histone modification, chromatin remodelling, and disruption of non-coding RNA [235]. Interestingly, while the majority of expressed miRNAs decline in the brain of aged animal models and humans, a small subset of non-coding RNAs was found selectively upregulated [236,237]. Epigenetic changes are also crucial for the induction, progression, and maintenance of senescence. In senescent human fibroblasts, the accumulation of a distinct heterochromatic structure (senescence-associated heterochromatic foci or SAHF) has been reported, possibly as a consequence of persistent DNA damage, decreased histone chaperone protein production, and decreased histone biosynthesis [238,239]. Moreover, histone acetylation seems to influence this phenotype. Indeed, Histone deacetylase 4 (HDAC4) is downregulated in oncogene-induced senescence, leading to the upregulation of senescence-associated genes [240]. Chromatin remodelling events and histones modifications, moreover, can influence the expression of SASP components [241]. Thus, the manipulation of these mechanisms is a prime target in age- and senescence-delaying interventions.
Even if mitochondria have their own genome, most of their proteins are nuclear-encoded and they need to communicate with the surrounding cellular environment. For this reason, the bi-directional communication between these organelles and the nucleus is constant and can lead to epigenetic modifications. In fact, mitochondrial dysfunctions invoke a process known as mitochondria-to-nucleus retrograde response, able to regulate nuclear-encoded gene expression and cellular metabolism [242]. Mitochondria provide numerous co-substrates produced in the Krebs cycle that are required for epigenetic and transcriptional processes, such as histone modifications and chromatin remodelling [243]. Changes in metabolism and metabolites level during ageing and senescence can therefore contribute to epigenetic modifications. In addition, mitochondria act as redox sensors able to identify stressful conditions and react by shaping the chromatin to promote survival or trigger senescence [244]. As a consequence, amelioration of mitochondrial functions would have positive effects also on the epigenetic state of the cell. Promising results have already been shown in in vivo models such as C. elegans [107], where elevated levels of the Krebs cycle intermediate α-ketoglutarate induced DNA and histone demethylation via activation of two histone demethylases, JMJD3 and PHF8, resulting in the removal of repressive marks, the induction of UPRmt gene expression, and extended lifespan. Similarly, mitochondrial ROS activate the DNA-damage-sensing kinases, Tel1p and Rad35p, resulting in enhanced subtelomeric silencing via inactivation of Rph1p, a histone H3K36 demethylase of the jumonji family of enzymes [245]. This represents another process through which the maintenance of mitochondrial homeostasis and ROS production under a certain threshold can promote longevity.

13. Transplant of Younger Mitochondria: Another Weapon against Ageing

So far, we have focused on identifying alterations happening in mitochondria in aged and senescent cells and how they can be exploited as potential targets for therapies. In aged cells, these strategies aim to reduce the damages that progressively accumulated in the “old” organelles over the years and improve their functions and structure. The majority of these alterations are shared by senescent cells and the targeting of these pathways can help eliminate them, reduce the SASP, or prevent the development of a senescent phenotype in the first place, when it is driven by mitochondrial dysfunctions. Recently, a less canonical idea to improve aged cells’ condition and reduce senescent cells accumulation in ageing-associated diseases has emerged: obtain cells with “younger” mitochondria through mitochondrial transplantation.
Mitochondrial transplant (mtTP) consists in extracting “young” and healthy mitochondria, injecting them into a patient, and allowing them to be absorbed into the cells. While the development of this technique in a laboratory setting is still at its initial stage, mtTP between cells is not an alien concept and has even been observed naturally in vitro and in vivo. For example, astrocytes close to the site of a focally induced cerebral ischaemia can transfer mitochondria to neurons [246]. Additionally, mitochondria can be transferred between cells through tunnelling nanotubes, thin plasma membrane structures connecting cells and allowing intercellular transfers of organelles, various plasma membrane components, and cytoplasmic molecules [247,248], or packed in extracellular vesicles [249]. In a laboratory setting, mitochondrial transfer has been attempted in numerous ways: via cytosol transfer, mitochondrial injection in cells, or injection in the bloodstream of animal models [250]. Intra-arterial injection of viable and respiration-competent autologous mitochondria isolated from pigs’ skeletal muscle was successfully used in the treatment of acute kidney injury and cardiac ischemia/reperfusion [251,252]. Similar results were shown in the lungs, where mitochondria were delivered either intra-arterially in the pulmonary artery or by a nebulizer [253]. Pre-labelled mitochondria were found up to 4 weeks after injection in the studied tissues, but their maintained functionality was unclear [251]. More recent studies demonstrated that mitochondrial transplant significantly up-regulates the expression of the mitochondrial complex II subunit SDHB in the hippocampus of aged mice [254] and improves basal respiration and ATP production 48 h post-transplantation in rats’ cardiomyocytes [255]. The observed benefits on respiration, however, appeared lost a month after the treatment, suggesting a potential for this technique in the treatment of acute injuries more than chronic conditions. Further research, optimisation, and technological advances are therefore necessary to determine if mtTP will be applicable in the treatment of diseases in the future, including age-associated pathologies.

14. Conclusions

Taken together, the evidence presented in this review shows that mitochondria dysfunctions have a close relationship with ageing and cellular senescence. Several mitochondrial pathways have already been taken into consideration as potential therapeutic targets for ageing-associated diseases, and promising compounds have been developed. Future research will have to answer numerous open questions including: is it possible to completely restore mitochondrial function in senescent and aged cells? Which age- or senescence-associated aspects are the primary drivers of mitochondrial dysfunction and vice-versa? Which ones are targetable therapeutically? Answering some of these questions could get us closer to healthy ageing, with countless medical, social and economic benefits.

Funding

This work was founded by European Commission under ‘Marie Skłodowska-Curie Actions’, Individual Fellowship (MitoSen-101060172) (M.P.). Work at the M.S. laboratory is funded by core grants from the IRB and “laCaixa” Foundation, and grants from the Spanish Ministry of Science co-funded by the European Regional Development Fund (ERDF) (SAF2017-82613-R), European Research Council (ERC-2014-AdG/669622), and Secretaria d’Universitats i Recerca del Departament d’Empresa i Coneixement of Catalonia (Grup de Recerca consolidat 2017 SGR 282).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

M.S. is shareholder of Senolytic Therapeutics, Life Biosciences, Rejuveron Senescence Therapeutics and Altos Labs, and is advisor of Rejuveron Senescence Therapeutics and Altos Labs. The funders had no role in the study design, decision to publish, or manuscript preparation.

References

  1. López-Otín, C.; Blasco, M.A.; Partridge, L.; Serrano, M.; Kroemer, G. The Hallmarks of Aging. Cell 2013, 153, 1194–1217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Hanahan, D. Hallmarks of Cancer: New Dimensions. Cancer Discov. 2022, 12, 31–46. [Google Scholar] [CrossRef] [PubMed]
  3. McHugh, D.; Gil, J. Senescence and Aging: Causes, Consequences, and Therapeutic Avenues. J. Cell Biol. 2018, 217, 65–77. [Google Scholar] [CrossRef] [PubMed]
  4. Gorgoulis, V.; Adams, P.D.; Alimonti, A.; Bennett, D.C.; Bischof, O.; Bishop, C.; Campisi, J.; Collado, M.; Evangelou, K.; Ferbeyre, G.; et al. Cellular Senescence: Defining a Path Forward. Cell 2019, 179, 813–827. [Google Scholar] [CrossRef] [PubMed]
  5. van Deursen, J.M. The Role of Senescent Cells in Ageing. Nature 2014, 509, 439–446. [Google Scholar] [CrossRef] [Green Version]
  6. Shen-Orr, S.S.; Furman, D.; Kidd, B.A.; Hadad, F.; Lovelace, P.; Huang, Y.-W.; Rosenberg-Hasson, Y.; Mackey, S.; Grisar, F.A.G.; Pickman, Y.; et al. Defective Signaling in the JAK-STAT Pathway Tracks with Chronic Inflammation and Cardiovascular Risk in Aging Humans. Cell Syst. 2016, 3, 374–384.e4. [Google Scholar] [CrossRef] [Green Version]
  7. De Nardo, D.; Latz, E. NLRP3 Inflammasomes Link Inflammation and Metabolic Disease. Trends Immunol. 2011, 32, 373–379. [Google Scholar] [CrossRef] [Green Version]
  8. Duewell, P.; Kono, H.; Rayner, K.J.; Sirois, C.M.; Vladimer, G.; Bauernfeind, F.G.; Abela, G.S.; Franchi, L.; Nuñez, G.; Schnurr, M.; et al. NLRP3 Inflammasomes Are Required for Atherogenesis and Activated by Cholesterol Crystals. Nature 2010, 464, 1357–1361. [Google Scholar] [CrossRef] [Green Version]
  9. Saresella, M.; La Rosa, F.; Piancone, F.; Zoppis, M.; Marventano, I.; Calabrese, E.; Rainone, V.; Nemni, R.; Mancuso, R.; Clerici, M. The NLRP3 and NLRP1 Inflammasomes Are Activated in Alzheimer’s Disease. Mol. Neurodegener. 2016, 11, 23. [Google Scholar] [CrossRef] [Green Version]
  10. Pansarasa, O.; Pistono, C.; Davin, A.; Bordoni, M.; Mimmi, M.C.; Guaita, A.; Cereda, C. Altered Immune System in Frailty: Genetics and Diet May Influence Inflammation. Ageing Res. Rev. 2019, 54, 100935. [Google Scholar] [CrossRef]
  11. Jeon, O.H.; Mehdipour, M.; Gil, T.-H.; Kang, M.; Aguirre, N.W.; Robinson, Z.R.; Kato, C.; Etienne, J.; Lee, H.G.; Alimirah, F.; et al. Systemic Induction of Senescence in Young Mice after Single Heterochronic Blood Exchange. Nat. Metab. 2022, 4, 995–1006. [Google Scholar] [CrossRef] [PubMed]
  12. Justice, J.N.; Nambiar, A.M.; Tchkonia, T.; LeBrasseur, N.K.; Pascual, R.; Hashmi, S.K.; Prata, L.; Masternak, M.M.; Kritchevsky, S.B.; Musi, N.; et al. Senolytics in Idiopathic Pulmonary Fibrosis: Results from a First-in-Human, Open-Label, Pilot Study. EBioMedicine 2019, 40, 554–563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Hickson, L.J.; Langhi Prata, L.G.P.; Bobart, S.A.; Evans, T.K.; Giorgadze, N.; Hashmi, S.K.; Herrmann, S.M.; Jensen, M.D.; Jia, Q.; Jordan, K.L.; et al. Senolytics Decrease Senescent Cells in Humans: Preliminary Report from a Clinical Trial of Dasatinib plus Quercetin in Individuals with Diabetic Kidney Disease. EBioMedicine 2019, 47, 446–456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Wiley, C.D.; Velarde, M.C.; Lecot, P.; Liu, S.; Sarnoski, E.A.; Freund, A.; Shirakawa, K.; Lim, H.W.; Davis, S.S.; Ramanathan, A.; et al. Mitochondrial Dysfunction Induces Senescence with a Distinct Secretory Phenotype. Cell Metab. 2016, 23, 303–314. [Google Scholar] [CrossRef] [Green Version]
  15. Correia-Melo, C.; Marques, F.D.M.; 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]
  16. Seong, S.-Y.; Matzinger, P. Hydrophobicity: An Ancient Damage-Associated Molecular Pattern That Initiates Innate Immune Responses. Nat. Rev. Immunol. 2004, 4, 469–478. [Google Scholar] [CrossRef]
  17. Dela Cruz, C.S.; Kang, M.-J. Mitochondrial Dysfunction and Damage Associated Molecular Patterns (DAMPs) in Chronic Inflammatory Diseases. Mitochondrion 2018, 41, 37–44. [Google Scholar] [CrossRef]
  18. Dorward, D.A.; Lucas, C.D.; Chapman, G.B.; Haslett, C.; Dhaliwal, K.; Rossi, A.G. The Role of Formylated Peptides and Formyl Peptide Receptor 1 in Governing Neutrophil Function during Acute Inflammation. Am. J. Pathol. 2015, 185, 1172–1184. [Google Scholar] [CrossRef] [Green Version]
  19. Tigano, M.; Vargas, D.C.; Tremblay-Belzile, S.; Fu, Y.; Sfeir, A. Nuclear Sensing of Breaks in Mitochondrial DNA Enhances Immune Surveillance. Nature 2021, 591, 477–481. [Google Scholar] [CrossRef]
  20. Busch, L.; Vieten, S.; Brödel, S.; Endres, K.; Bufe, B. Emerging Contributions of Formyl Peptide Receptors to Neurodegenerative Diseases. Biol. Chem. 2022, 403, 27–41. [Google Scholar] [CrossRef]
  21. Schröder, N.; Schaffrath, A.; Welter, J.A.; Putzka, T.; Griep, A.; Ziegler, P.; Brandt, E.; Samer, S.; Heneka, M.T.; Kaddatz, H.; et al. Inhibition of Formyl Peptide Receptors Improves the Outcome in a Mouse Model of Alzheimer Disease. J. Neuroinflammation 2020, 17, 131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Chinnery, P.F.; Hudson, G. Mitochondrial Genetics. Br. Med. Bull. 2013, 106, 135–159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Wolf, A.M. MtDNA Mutations and Aging—Not a Closed Case after All? Sig. Transduct. Target. Ther. 2021, 6, 1–2. [Google Scholar] [CrossRef] [PubMed]
  24. Lee, H.-C.; Wei, Y.-H. Mitochondria and Aging. Adv. Exp. Med. Biol. 2012, 942, 311–327. [Google Scholar] [CrossRef] [PubMed]
  25. Fayet, G.; Jansson, M.; Sternberg, D.; Moslemi, A.R.; Blondy, P.; Lombès, A.; Fardeau, M.; Oldfors, A. Ageing Muscle: Clonal Expansions of Mitochondrial DNA Point Mutations and Deletions Cause Focal Impairment of Mitochondrial Function. Neuromuscul. Disord. 2002, 12, 484–493. [Google Scholar] [CrossRef] [PubMed]
  26. Eshaghian, A.; Vleugels, R.A.; Canter, J.A.; McDonald, M.A.; Stasko, T.; Sligh, J.E. Mitochondrial DNA Deletions Serve as Biomarkers of Aging in the Skin, but Are Typically Absent in Nonmelanoma Skin Cancers. J. Investig. Dermatol. 2006, 126, 336–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Trifunovic, A.; Wredenberg, A.; Falkenberg, M.; Spelbrink, J.N.; Rovio, A.T.; Bruder, C.E.; Bohlooly-Y, M.; Gidlöf, S.; Oldfors, A.; Wibom, R.; et al. Premature Ageing in Mice Expressing Defective Mitochondrial DNA Polymerase. Nature 2004, 429, 417–423. [Google Scholar] [CrossRef]
  28. Kujoth, G.C.; Hiona, A.; Pugh, T.D.; Someya, S.; Panzer, K.; Wohlgemuth, S.E.; Hofer, T.; Seo, A.Y.; Sullivan, R.; Jobling, W.A.; et al. Mitochondrial DNA Mutations, Oxidative Stress, and Apoptosis in Mammalian Aging. Science 2005, 309, 481–484. [Google Scholar] [CrossRef]
  29. Vermulst, M.; Bielas, J.H.; Kujoth, G.C.; Ladiges, W.C.; Rabinovitch, P.S.; Prolla, T.A.; Loeb, L.A. Mitochondrial Point Mutations Do Not Limit the Natural Lifespan of Mice. Nat. Genet. 2007, 39, 540–543. [Google Scholar] [CrossRef]
  30. Smith, A.L.; Whitehall, J.C.; Bradshaw, C.; Gay, D.; Robertson, F.; Blain, A.P.; Hudson, G.; Pyle, A.; Houghton, D.; Hunt, M.; et al. Age-Associated Mitochondrial DNA Mutations Cause Metabolic Remodelling That Contributes to Accelerated Intestinal Tumorigenesis. Nat. Cancer 2020, 1, 976–989. [Google Scholar] [CrossRef]
  31. Yuan, Y.; Ju, Y.S.; Kim, Y.; Li, J.; Wang, Y.; Yoon, C.J.; Yang, Y.; Martincorena, I.; Creighton, C.J.; Weinstein, J.N.; et al. Comprehensive Molecular Characterization of Mitochondrial Genomes in Human Cancers. Nat. Genet. 2020, 52, 342–352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Sub-Lethal Apoptotic Stress Enables MtDNA Release during Senescence and Drives the SASP. Available online: https://www.researchsquare.com (accessed on 20 October 2022).
  33. Decout, A.; Katz, J.D.; Venkatraman, S.; Ablasser, A. The CGAS-STING Pathway as a Therapeutic Target in Inflammatory Diseases. Nat. Rev. Immunol. 2021, 21, 548–569. [Google Scholar] [CrossRef] [PubMed]
  34. Hopfner, K.-P.; Hornung, V. Molecular Mechanisms and Cellular Functions of CGAS-STING Signalling. Nat. Rev. Mol. Cell Biol. 2020, 21, 501–521. [Google Scholar] [CrossRef] [PubMed]
  35. Yu, C.-H.; Davidson, S.; Harapas, C.R.; Hilton, J.B.; Mlodzianoski, M.J.; Laohamonthonkul, P.; Louis, C.; Low, R.R.J.; Moecking, J.; De Nardo, D.; et al. TDP-43 Triggers Mitochondrial DNA Release via MPTP to Activate CGAS/STING in ALS. Cell 2020, 183, 636–649.e18. [Google Scholar] [CrossRef]
  36. Chung, K.W.; Dhillon, P.; Huang, S.; Sheng, X.; Shrestha, R.; Qiu, C.; Kaufman, B.A.; Park, J.; Pei, L.; Baur, J.; et al. Mitochondrial Damage and Activation of the STING Pathway Lead to Renal Inflammation and Fibrosis. Cell Metab. 2019, 30, 784–799.e5. [Google Scholar] [CrossRef]
  37. Huang, L.S.; Hong, Z.; Wu, W.; Xiong, S.; Zhong, M.; Gao, X.; Rehman, J.; Malik, A.B. MtDNA Activates CGAS Signaling and Suppresses the YAP-Mediated Endothelial Cell Proliferation Program to Promote Inflammatory Injury. Immunity 2020, 52, 475–486.e5. [Google Scholar] [CrossRef]
  38. Jauhari, A.; Baranov, S.V.; Suofu, Y.; Kim, J.; Singh, T.; Yablonska, S.; Li, F.; Wang, X.; Oberly, P.; Minnigh, M.B.; et al. Melatonin Inhibits Cytosolic Mitochondrial DNA–Induced Neuroinflammatory Signaling in Accelerated Aging and Neurodegeneration. J. Clin. Investig. 2020, 130, 3124–3136. [Google Scholar] [CrossRef] [Green Version]
  39. Carroll Elizabeth, C.; Jin, L.; Mori, A.; Muñoz-Wolf, N.; Oleszycka, E.; Moran, H.B.T.; Mansouri, S.; McEntee, C.P.; Lambe, E.; Agger, E.M.; et al. The Vaccine Adjuvant Chitosan Promotes Cellular Immunity via DNA Sensor CGAS-STING-Dependent Induction of Type I Interferons. Immunity 2016, 44, 597–608. [Google Scholar] [CrossRef] [Green Version]
  40. Vincent, J.; Adura, C.; Gao, P.; Luz, A.; Lama, L.; Asano, Y.; Okamoto, R.; Imaeda, T.; Aida, J.; Rothamel, K.; et al. Small Molecule Inhibition of CGAS Reduces Interferon Expression in Primary Macrophages from Autoimmune Mice. Nat. Commun. 2017, 8, 750. [Google Scholar] [CrossRef] [Green Version]
  41. Padilla-Salinas, R.; Sun, L.; Anderson, R.; Yang, X.; Zhang, S.; Chen, Z.J.; Yin, H. Discovery of Small-Molecule Cyclic GMP-AMP Synthase Inhibitors. J. Org. Chem. 2020, 85, 1579–1600. [Google Scholar] [CrossRef]
  42. Lama, L.; Adura, C.; Xie, W.; Tomita, D.; Kamei, T.; Kuryavyi, V.; Gogakos, T.; Steinberg, J.I.; Miller, M.; Ramos-Espiritu, L.; et al. Development of Human CGAS-Specific Small-Molecule Inhibitors for Repression of DsDNA-Triggered Interferon Expression. Nat. Commun. 2019, 10, 2261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Hall, J.; Brault, A.; Vincent, F.; Weng, S.; Wang, H.; Dumlao, D.; Aulabaugh, A.; Aivazian, D.; Castro, D.; Chen, M.; et al. Discovery of PF-06928215 as a High Affinity Inhibitor of CGAS Enabled by a Novel Fluorescence Polarization Assay. PLoS ONE 2017, 12, e0184843. [Google Scholar] [CrossRef]
  44. Bauer, S.; Kirschning, C.J.; Häcker, H.; Redecke, V.; Hausmann, S.; Akira, S.; Wagner, H.; Lipford, G.B. Human TLR9 Confers Responsiveness to Bacterial DNA via Species-Specific CpG Motif Recognition. Proc. Natl. Acad. Sci. USA 2001, 98, 9237–9242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Krug, A.; French, A.R.; Barchet, W.; Fischer, J.A.A.; Dzionek, A.; Pingel, J.T.; Orihuela, M.M.; Akira, S.; Yokoyama, W.M.; Colonna, M. TLR9-Dependent Recognition of MCMV by IPC and DC Generates Coordinated Cytokine Responses That Activate Antiviral NK Cell Function. Immunity 2004, 21, 107–119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. West, A.P.; Shadel, G.S. Mitochondrial DNA in Innate Immune Responses and Inflammatory Pathology. Nat. Rev. Immunol. 2017, 17, 363–375. [Google Scholar] [CrossRef]
  47. Hoque, R.; Sohail, M.; Malik, A.; Sarwar, S.; Luo, Y.; Shah, A.; Barrat, F.; Flavell, R.; Gorelick, F.; Husain, S.; et al. TLR9 and the NLRP3 Inflammasome Link Acinar Cell Death with Inflammation in Acute Pancreatitis. Gastroenterology 2011, 141, 358–369. [Google Scholar] [CrossRef] [Green Version]
  48. Pinti, M.; Cevenini, E.; Nasi, M.; De Biasi, S.; Salvioli, S.; Monti, D.; Benatti, S.; Gibellini, L.; Cotichini, R.; Stazi, M.A.; et al. Circulating Mitochondrial DNA Increases with Age and Is a Familiar Trait: Implications for “Inflamm-Aging”. Eur. J. Immunol. 2014, 44, 1552–1562. [Google Scholar] [CrossRef]
  49. Tilokani, L.; Nagashima, S.; Paupe, V.; Prudent, J. Mitochondrial Dynamics: Overview of Molecular Mechanisms. Essays Biochem. 2018, 62, 341–360. [Google Scholar] [CrossRef] [Green Version]
  50. Pickles, S.; Vigié, P.; Youle, R.J. Mitophagy and Quality Control Mechanisms in Mitochondrial Maintenance. Curr. Biol. 2018, 28, R170–R185. [Google Scholar] [CrossRef] [Green Version]
  51. Leduc-Gaudet, J.-P.; Picard, M.; St-Jean Pelletier, F.; Sgarioto, N.; Auger, M.-J.; Vallée, J.; Robitaille, R.; St-Pierre, D.H.; Gouspillou, G. Mitochondrial Morphology Is Altered in Atrophied Skeletal Muscle of Aged Mice. Oncotarget 2015, 6, 17923–17937. [Google Scholar] [CrossRef]
  52. Terman, A.; Brunk, U.T. The Aging Myocardium: Roles of Mitochondrial Damage and Lysosomal Degradation. Heart Lung Circ. 2005, 14, 107–114. [Google Scholar] [CrossRef]
  53. 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]
  54. 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] [PubMed]
  55. 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] [PubMed] [Green Version]
  56. Youle, R.J.; van der Bliek, A.M. Mitochondrial Fission, Fusion, and Stress. Science 2012, 337, 1062–1065. [Google Scholar] [CrossRef] [Green Version]
  57. Estaquier, J.; Arnoult, D. Inhibiting Drp1-Mediated Mitochondrial Fission Selectively Prevents the Release of Cytochrome c during Apoptosis. Cell Death Differ. 2007, 14, 1086–1094. [Google Scholar] [CrossRef] [Green Version]
  58. Frank, S.; Gaume, B.; Bergmann-Leitner, E.S.; Leitner, W.W.; Robert, E.G.; Catez, F.; Smith, C.L.; Youle, R.J. The Role of Dynamin-Related Protein 1, a Mediator of Mitochondrial Fission, in Apoptosis. Dev. Cell 2001, 1, 515–525. [Google Scholar] [CrossRef] [Green Version]
  59. Yu, B.; Ma, J.; Li, J.; Wang, D.; Wang, Z.; Wang, S. Mitochondrial Phosphatase PGAM5 Modulates Cellular Senescence by Regulating Mitochondrial Dynamics. Nat. Commun. 2020, 11, 2549. [Google Scholar] [CrossRef]
  60. Rana, A.; Oliveira, M.P.; Khamoui, A.V.; Aparicio, R.; Rera, M.; Rossiter, H.B.; Walker, D.W. Promoting Drp1-Mediated Mitochondrial Fission in Midlife Prolongs Healthy Lifespan of Drosophila Melanogaster. Nat. Commun. 2017, 8, 448. [Google Scholar] [CrossRef] [Green Version]
  61. D’Amico, D.; Mottis, A.; Potenza, F.; Sorrentino, V.; Li, H.; Romani, M.; Lemos, V.; Schoonjans, K.; Zamboni, N.; Knott, G.; et al. The RNA-Binding Protein PUM2 Impairs Mitochondrial Dynamics and Mitophagy During Aging. Mol. Cell 2019, 73, 775–787.e10. [Google Scholar] [CrossRef]
  62. Korolchuk, V.I.; Miwa, S.; Carroll, B.; von Zglinicki, T. Mitochondria in Cell Senescence: Is Mitophagy the Weakest Link? EBioMedicine 2017, 21, 7–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Chen, G.; Kroemer, G.; Kepp, O. Mitophagy: An Emerging Role in Aging and Age-Associated Diseases. Front. Cell Dev. Biol. 2020, 8, 200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Gouspillou, G.; Sgarioto, N.; Kapchinsky, S.; Purves-Smith, F.; Norris, B.; Pion, C.H.; Barbat-Artigas, S.; Lemieux, F.; Taivassalo, T.; Morais, J.A.; et al. Increased Sensitivity to Mitochondrial Permeability Transition and Myonuclear Translocation of Endonuclease G in Atrophied Muscle of Physically Active Older Humans. FASEB J. 2014, 28, 1621–1633. [Google Scholar] [CrossRef]
  65. O’Leary, M.F.; Vainshtein, A.; Iqbal, S.; Ostojic, O.; Hood, D.A. Adaptive Plasticity of Autophagic Proteins to Denervation in Aging Skeletal Muscle. Am. J. Physiol. Cell Physiol. 2013, 304, C422–C430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Dalle Pezze, P.; Nelson, G.; Otten, E.G.; Korolchuk, V.I.; Kirkwood, T.B.L.; von Zglinicki, T.; Shanley, D.P. Dynamic Modelling of Pathways to Cellular Senescence Reveals Strategies for Targeted Interventions. PLoS Comput. Biol. 2014, 10, e1003728. [Google Scholar] [CrossRef] [Green Version]
  67. García-Prat, L.; Martínez-Vicente, M.; Perdiguero, E.; Ortet, L.; Rodríguez-Ubreva, J.; Rebollo, E.; Ruiz-Bonilla, V.; Gutarra, S.; Ballestar, E.; Serrano, A.L.; et al. Autophagy Maintains Stemness by Preventing Senescence. Nature 2016, 529, 37–42. [Google Scholar] [CrossRef]
  68. Kissová, I.; Deffieu, M.; Manon, S.; Camougrand, N. Uth1p Is Involved in the Autophagic Degradation of Mitochondria. J. Biol. Chem. 2004, 279, 39068–39074. [Google Scholar] [CrossRef] [Green Version]
  69. Sitte, N.; Merker, K.; Grune, T.; von Zglinicki, T. Lipofuscin Accumulation in Proliferating Fibroblasts in Vitro: An Indicator of Oxidative Stress. Exp. Gerontol. 2001, 36, 475–486. [Google Scholar] [CrossRef]
  70. Demidenko, Z.N.; Blagosklonny, M.V. Growth Stimulation Leads to Cellular Senescence When the Cell Cycle Is Blocked. Cell Cycle 2008, 7, 3355–3361. [Google Scholar] [CrossRef]
  71. Carroll, B.; Nelson, G.; Rabanal-Ruiz, Y.; Kucheryavenko, O.; Dunhill-Turner, N.A.; Chesterman, C.C.; Zahari, Q.; Zhang, T.; Conduit, S.E.; Mitchell, C.A.; et al. Persistent MTORC1 Signaling in Cell Senescence Results from Defects in Amino Acid and Growth Factor Sensing. J. Cell Biol. 2017, 216, 1949–1957. [Google Scholar] [CrossRef]
  72. Blagosklonny, M.V. Rapamycin for Longevity: Opinion Article. Aging 2019, 11, 8048–8067. [Google Scholar] [CrossRef] [PubMed]
  73. Schinaman, J.M.; Rana, A.; Ja, W.W.; Clark, R.I.; Walker, D.W. Rapamycin Modulates Tissue Aging and Lifespan Independently of the Gut Microbiota in Drosophila. Sci. Rep. 2019, 9, 7824. [Google Scholar] [CrossRef] [Green Version]
  74. Juricic, P.; Lu, Y.-X.; Leech, T.; Drews, L.F.; Paulitz, J.; Lu, J.; Nespital, T.; Azami, S.; Regan, J.C.; Funk, E.; et al. Long-Lasting Geroprotection from Brief Rapamycin Treatment in Early Adulthood by Persistently Increased Intestinal Autophagy. Nat. Aging 2022, 2, 824–836. [Google Scholar] [CrossRef]
  75. Bordi, M.; Darji, S.; Sato, Y.; Mellén, M.; Berg, M.J.; Kumar, A.; Jiang, Y.; Nixon, R.A. MTOR Hyperactivation in Down Syndrome Underlies Deficits in Autophagy Induction, Autophagosome Formation, and Mitophagy. Cell Death Dis. 2019, 10, 563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Twig, G.; Shirihai, O.S. The Interplay between Mitochondrial Dynamics and Mitophagy. Antioxid. Redox Signal. 2011, 14, 1939–1951. [Google Scholar] [CrossRef] [Green Version]
  77. Bueno, M.; Lai, Y.-C.; Romero, Y.; Brands, J.; St Croix, C.M.; Kamga, C.; Corey, C.; Herazo-Maya, J.D.; Sembrat, J.; Lee, J.S.; et al. PINK1 Deficiency Impairs Mitochondrial Homeostasis and Promotes Lung Fibrosis. J. Clin. Investig. 2015, 125, 521–538. [Google Scholar] [CrossRef] [Green Version]
  78. Narendra, D.; Tanaka, A.; Suen, D.-F.; Youle, R.J. Parkin Is Recruited Selectively to Impaired Mitochondria and Promotes Their Autophagy. J. Cell Biol. 2008, 183, 795–803. [Google Scholar] [CrossRef] [Green Version]
  79. Chistiakov, D.A.; Sobenin, I.A.; Revin, V.V.; Orekhov, A.N.; Bobryshev, Y.V. Mitochondrial Aging and Age-Related Dysfunction of Mitochondria. Biomed. Res. Int. 2014, 2014, 238463. [Google Scholar] [CrossRef] [Green Version]
  80. Summer, R.; Shaghaghi, H.; Schriner, D.; Roque, W.; Sales, D.; Cuevas-Mora, K.; Desai, V.; Bhushan, A.; Ramirez, M.I.; Romero, F. Activation of the MTORC1/PGC-1 Axis Promotes Mitochondrial Biogenesis and Induces Cellular Senescence in the Lung Epithelium. Am. J. Physiol. Lung Cell. Mol. Physiol. 2019, 316, L1049–L1060. [Google Scholar] [CrossRef]
  81. Sczelecki, S.; Besse-Patin, A.; Abboud, A.; Kleiner, S.; Laznik-Bogoslavski, D.; Wrann, C.D.; Ruas, J.L.; Haibe-Kains, B.; Estall, J.L. Loss of Pgc-1α Expression in Aging Mouse Muscle Potentiates Glucose Intolerance and Systemic Inflammation. Am. J. Physiol. Endocrinol. Metab. 2014, 306, E157–E167. [Google Scholar] [CrossRef]
  82. Arany, Z.; Wagner, B.K.; Ma, Y.; Chinsomboon, J.; Laznik, D.; Spiegelman, B.M. Gene Expression-Based Screening Identifies Microtubule Inhibitors as Inducers of PGC-1alpha and Oxidative Phosphorylation. Proc. Natl. Acad. Sci. USA 2008, 105, 4721–4726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Tan, Z.; Luo, X.; Xiao, L.; Tang, M.; Bode, A.M.; Dong, Z.; Cao, Y. The Role of PGC1α in Cancer Metabolism and Its Therapeutic Implications. Mol. Cancer Ther. 2016, 15, 774–782. [Google Scholar] [CrossRef] [Green Version]
  84. Reznick, R.M.; Zong, H.; Li, J.; Morino, K.; Moore, I.K.; Yu, H.J.; Liu, Z.-X.; Dong, J.; Mustard, K.J.; Hawley, S.A.; et al. Aging-Associated Reductions in AMP-Activated Protein Kinase Activity and Mitochondrial Biogenesis. Cell Metab. 2007, 5, 151–156. [Google Scholar] [CrossRef] [Green Version]
  85. Qiang, W.; Weiqiang, K.; Qing, Z.; Pengju, Z.; Yi, L. Aging Impairs Insulin-Stimulated Glucose Uptake in Rat Skeletal Muscle via Suppressing AMPKalpha. Exp. Mol. Med. 2007, 39, 535–543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Curtis, R.; O’Connor, G.; DiStefano, P.S. Aging Networks in Caenorhabditis Elegans: AMP-Activated Protein Kinase (Aak-2) Links Multiple Aging and Metabolism Pathways. Aging Cell 2006, 5, 119–126. [Google Scholar] [CrossRef]
  87. McCarty, M.F. Chronic Activation of AMP-Activated Kinase as a Strategy for Slowing Aging. Med. Hypotheses 2004, 63, 334–339. [Google Scholar] [CrossRef]
  88. Mohammed, I.; Hollenberg, M.D.; Ding, H.; Triggle, C.R. A Critical Review of the Evidence That Metformin Is a Putative Anti-Aging Drug That Enhances Healthspan and Extends Lifespan. Front. Endocrinol. 2021, 12, 718942. [Google Scholar] [CrossRef] [PubMed]
  89. Zhou, J.; Massey, S.; Story, D.; Li, L. Metformin: An Old Drug with New Applications. Int. J. Mol. Sci. 2018, 19, 2863. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Triggle, C.R.; Mohammed, I.; Bshesh, K.; Marei, I.; Ye, K.; Ding, H.; MacDonald, R.; Hollenberg, M.D.; Hill, M.A. Metformin: Is It a Drug for All Reasons and Diseases? Metabolism 2022, 133, 155223. [Google Scholar] [CrossRef]
  91. Martin-Montalvo, A.; Mercken, E.M.; Mitchell, S.J.; Palacios, H.H.; Mote, P.L.; Scheibye-Knudsen, M.; Gomes, A.P.; Ward, T.M.; Minor, R.K.; Blouin, M.-J.; et al. Metformin Improves Healthspan and Lifespan in Mice. Nat. Commun. 2013, 4, 2192. [Google Scholar] [CrossRef]
  92. Strong, R.; Miller, R.A.; Antebi, A.; Astle, C.M.; Bogue, M.; Denzel, M.S.; Fernandez, E.; Flurkey, K.; Hamilton, K.L.; Lamming, D.W.; et al. Longer Lifespan in Male Mice Treated with a Weakly Estrogenic Agonist, an Antioxidant, an α-Glucosidase Inhibitor or a Nrf2-Inducer. Aging Cell 2016, 15, 872–884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Zhu, X.; Shen, W.; Liu, Z.; Sheng, S.; Xiong, W.; He, R.; Zhang, X.; Ma, L.; Ju, Z. Effect of Metformin on Cardiac Metabolism and Longevity in Aged Female Mice. Front. Cell Dev. Biol. 2020, 8, 626011. [Google Scholar] [CrossRef] [PubMed]
  94. Wang, W.; Yang, X.; López de Silanes, I.; Carling, D.; Gorospe, M. Increased AMP:ATP Ratio and AMP-Activated Protein Kinase Activity during Cellular Senescence Linked to Reduced HuR Function. J. Biol. Chem. 2003, 278, 27016–27023. [Google Scholar] [CrossRef] [Green Version]
  95. Han, X.; Tai, H.; Wang, X.; Wang, Z.; Zhou, J.; Wei, X.; Ding, Y.; Gong, H.; Mo, C.; Zhang, J.; et al. AMPK Activation Protects Cells from Oxidative Stress-Induced Senescence via Autophagic Flux Restoration and Intracellular NAD(+) Elevation. Aging Cell 2016, 15, 416–427. [Google Scholar] [CrossRef] [Green Version]
  96. Owusu-Ansah, E.; Song, W.; Perrimon, N. Muscle Mitohormesis Promotes Longevity via Systemic Repression of Insulin Signaling. Cell 2013, 155, 699–712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Yee, C.; Yang, W.; Hekimi, S. The Intrinsic Apoptosis Pathway Mediates the Pro-Longevity Response to Mitochondrial ROS in C. Elegans. Cell 2014, 157, 897–909. [Google Scholar] [CrossRef] [Green Version]
  98. Shpilka, T.; Haynes, C.M. The Mitochondrial UPR: Mechanisms, Physiological Functions and Implications in Ageing. Nat. Rev. Mol. Cell Biol. 2018, 19, 109–120. [Google Scholar] [CrossRef]
  99. Schulz, A.M.; Haynes, C.M. UPR(Mt)-Mediated Cytoprotection and Organismal Aging. Biochim. Biophys. Acta 2015, 1847, 1448–1456. [Google Scholar] [CrossRef] [Green Version]
  100. Nargund, A.M.; Fiorese, C.J.; Pellegrino, M.W.; Deng, P.; Haynes, C.M. Mitochondrial and Nuclear Accumulation of the Transcription Factor ATFS-1 Promotes OXPHOS Recovery during the UPR(Mt). Mol. Cell 2015, 58, 123–133. [Google Scholar] [CrossRef] [Green Version]
  101. Naresh, N.U.; Haynes, C.M. Signaling and Regulation of the Mitochondrial Unfolded Protein Response. Cold Spring Harb. Perspect. Biol. 2019, 11, a033944. [Google Scholar] [CrossRef]
  102. Houtkooper, R.H.; Mouchiroud, L.; Ryu, D.; Moullan, N.; Katsyuba, E.; Knott, G.; Williams, R.W.; Auwerx, J. Mitonuclear Protein Imbalance as a Conserved Longevity Mechanism. Nature 2013, 497, 451–457. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Durieux, J.; Wolff, S.; Dillin, A. The Cell-Non-Autonomous Nature of Electron Transport Chain-Mediated Longevity. Cell 2011, 144, 79–91. [Google Scholar] [CrossRef] [Green Version]
  104. Dillin, A.; Hsu, A.-L.; Arantes-Oliveira, N.; Lehrer-Graiwer, J.; Hsin, H.; Fraser, A.G.; Kamath, R.S.; Ahringer, J.; Kenyon, C. Rates of Behavior and Aging Specified by Mitochondrial Function during Development. Science 2002, 298, 2398–2401. [Google Scholar] [CrossRef]
  105. Borch Jensen, M.; Qi, Y.; Riley, R.; Rabkina, L.; Jasper, H. PGAM5 Promotes Lasting FoxO Activation after Developmental Mitochondrial Stress and Extends Lifespan in Drosophila. eLife 2017, 6, e26952. [Google Scholar] [CrossRef] [PubMed]
  106. Tian, Y.; Garcia, G.; Bian, Q.; Steffen, K.K.; Joe, L.; Wolff, S.; Meyer, B.J.; Dillin, A. Mitochondrial Stress Induces Chromatin Reorganization to Promote Longevity and UPR(Mt). Cell 2016, 165, 1197–1208. [Google Scholar] [CrossRef] [Green Version]
  107. Merkwirth, C.; Jovaisaite, V.; Durieux, J.; Matilainen, O.; Jordan, S.D.; Quiros, P.M.; Steffen, K.K.; Williams, E.G.; Mouchiroud, L.; Tronnes, S.U.; et al. Two Conserved Histone Demethylases Regulate Mitochondrial Stress-Induced Longevity. Cell 2016, 165, 1209–1223. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Sen, B.; Rastogi, A.; Nath, R.; Shasthry, S.M.; Pamecha, V.; Pandey, S.; Gupta, K.J.; Sarin, S.K.; Trehanpati, N.; Ramakrishna, G. Senescent Hepatocytes in Decompensated Liver Show Reduced UPRMT and Its Key Player, CLPP, Attenuates Senescence In Vitro. Cell Mol. Gastroenterol. Hepatol. 2019, 8, 73–94. [Google Scholar] [CrossRef] [Green Version]
  109. Lee, T.-Y.; Huang, L.-J.; Dong, H.-P.; Tohru, Y.; Liu, B.-H.; Yang, R.-C. Impairment of Mitochondrial Unfolded Protein Response Contribute to Resistance Declination of H2 O2 -Induced Injury in Senescent MRC-5 Cell Model. Kaohsiung J. Med. Sci. 2020, 36, 89–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Protasoni, M.; Zeviani, M. Mitochondrial Structure and Bioenergetics in Normal and Disease Conditions. Int. J. Mol. Sci. 2021, 22, 586. [Google Scholar] [CrossRef]
  111. Conley, K.E.; Jubrias, S.A.; Esselman, P.C. Oxidative Capacity and Ageing in Human Muscle. J. Physiol. 2000, 526 Pt 1, 203–210. [Google Scholar] [CrossRef]
  112. Santanasto, A.J.; Glynn, N.W.; Jubrias, S.A.; Conley, K.E.; Boudreau, R.M.; Amati, F.; Mackey, D.C.; Simonsick, E.M.; Strotmeyer, E.S.; Coen, P.M.; et al. Skeletal Muscle Mitochondrial Function and Fatigability in Older Adults. J. Gerontol. A Biol. Sci. Med. Sci. 2015, 70, 1379–1385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Frasca, D.; Saada, Y.B.; Garcia, D.; Friguet, B. Effects of Cellular Senescence on Metabolic Pathways in Non-Immune and Immune Cells. Mech. Ageing Dev. 2021, 194, 111428. [Google Scholar] [CrossRef] [PubMed]
  114. Kaplon, J.; Zheng, L.; Meissl, K.; Chaneton, B.; Selivanov, V.A.; Mackay, G.; van der Burg, S.H.; Verdegaal, E.M.E.; Cascante, M.; Shlomi, T.; et al. A Key Role for Mitochondrial Gatekeeper Pyruvate Dehydrogenase in Oncogene-Induced Senescence. Nature 2013, 498, 109–112. [Google Scholar] [CrossRef]
  115. Hutter, E.; Renner, K.; Pfister, G.; Stöckl, P.; Jansen-Dürr, P.; Gnaiger, E. Senescence-Associated Changes in Respiration and Oxidative Phosphorylation in Primary Human Fibroblasts. Biochem. J. 2004, 380, 919–928. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. 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]
  117. James, E.L.; Michalek, R.D.; Pitiyage, G.N.; de Castro, A.M.; Vignola, K.S.; Jones, J.; Mohney, R.P.; Karoly, E.D.; Prime, S.S.; Parkinson, E.K. Senescent Human Fibroblasts Show Increased Glycolysis and Redox Homeostasis with Extracellular Metabolomes That Overlap with Those of Irreparable DNA Damage, Aging, and Disease. J. Proteome Res. 2015, 14, 1854–1871. [Google Scholar] [CrossRef] [PubMed]
  118. Quijano, C.; Cao, L.; Fergusson, M.M.; Romero, H.; Liu, J.; Gutkind, S.; Rovira, I.I.; Mohney, R.P.; Karoly, E.D.; Finkel, T. Oncogene-Induced Senescence Results in Marked Metabolic and Bioenergetic Alterations. Cell Cycle 2012, 11, 1383–1392. [Google Scholar] [CrossRef] [Green Version]
  119. Marcheggiani, F.; Cirilli, I.; Orlando, P.; Silvestri, S.; Vogelsang, A.; Knott, A.; Blatt, T.; Weise, J.M.; Tiano, L. Modulation of Coenzyme Q10 Content and Oxidative Status in Human Dermal Fibroblasts Using HMG-CoA Reductase Inhibitor over a Broad Range of Concentrations. From Mitohormesis to Mitochondrial Dysfunction and Accelerated Aging. Aging 2019, 11, 2565–2582. [Google Scholar] [CrossRef]
  120. de Barcelos, I.P.; Haas, R.H. CoQ10 and Aging. Biology 2019, 8, 28. [Google Scholar] [CrossRef] [Green Version]
  121. Alcázar-Fabra, M.; Navas, P.; Brea-Calvo, G. Coenzyme Q Biosynthesis and Its Role in the Respiratory Chain Structure. Biochim. Biophys. Acta 2016, 1857, 1073–1078. [Google Scholar] [CrossRef]
  122. Tian, G.; Sawashita, J.; Kubo, H.; Nishio, S.; 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]
  123. Zhang, D.; Yan, B.; Yu, S.; Zhang, C.; Wang, B.; Wang, Y.; Wang, J.; Yuan, Z.; Zhang, L.; Pan, J. Coenzyme Q10 Inhibits the Aging of Mesenchymal Stem Cells Induced by D-Galactose through Akt/MTOR Signaling. Oxid. Med. Cell Longev. 2015, 2015, 867293. [Google Scholar] [CrossRef] [Green Version]
  124. Huo, J.; Xu, Z.; Hosoe, K.; Kubo, H.; Miyahara, H.; Dai, J.; Mori, M.; Sawashita, J.; Higuchi, K. Coenzyme Q10 Prevents Senescence and Dysfunction Caused by Oxidative Stress in Vascular Endothelial Cells. Oxid. Med. Cell Longev. 2018, 2018, 3181759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Champigny, C.M.; Cormier, R.P.J.; Simard, C.J.; St-Coeur, P.-D.; Fortin, S.; Pichaud, N. Omega-3 Monoacylglyceride Effects on Longevity, Mitochondrial Metabolism and Oxidative Stress: Insights from Drosophila Melanogaster. Mar. Drugs 2018, 16, 453. [Google Scholar] [CrossRef] [Green Version]
  126. Afshordel, S.; Hagl, S.; Werner, D.; Röhner, N.; Kögel, D.; Bazan, N.G.; Eckert, G.P. Omega-3 Polyunsaturated Fatty Acids Improve Mitochondrial Dysfunction in Brain Aging--Impact of Bcl-2 and NPD-1 like Metabolites. Prostaglandins Leukot. Essent. Fat. Acids 2015, 92, 23–31. [Google Scholar] [CrossRef] [PubMed]
  127. Lalia, A.Z.; Dasari, S.; Robinson, M.M.; Abid, H.; Morse, D.M.; Klaus, K.A.; Lanza, I.R. Influence of Omega-3 Fatty Acids on Skeletal Muscle Protein Metabolism and Mitochondrial Bioenergetics in Older Adults. Aging 2017, 9, 1096–1129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Mann, S.N.; Hadad, N.; Nelson Holte, M.; Rothman, A.R.; Sathiaseelan, R.; Ali Mondal, S.; Agbaga, M.-P.; Unnikrishnan, A.; Subramaniam, M.; Hawse, J.; et al. Health Benefits Attributed to 17α-Estradiol, a Lifespan-Extending Compound, Are Mediated through Estrogen Receptor α. eLife 2020, 9, e59616. [Google Scholar] [CrossRef]
  129. Biochemistry, Electron Transport Chain-StatPearls-NCBI Bookshelf. Available online: https://www.ncbi.nlm.nih.gov/books/NBK526105/ (accessed on 11 October 2022).
  130. Sies, H.; Berndt, C.; Jones, D.P. Oxidative Stress. Annu. Rev. Biochem. 2017, 86, 715–748. [Google Scholar] [CrossRef]
  131. Gelman, S.J.; Naser, F.; Mahieu, N.G.; McKenzie, L.D.; Dunn, G.P.; Chheda, M.G.; Patti, G.J. Consumption of NADPH for 2-HG Synthesis Increases Pentose Phosphate Pathway Flux and Sensitizes Cells to Oxidative Stress. Cell Rep. 2018, 22, 512–522. [Google Scholar] [CrossRef] [Green Version]
  132. Gomes, A.P.; Price, N.L.; Ling, A.J.Y.; 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]
  133. 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] [Green Version]
  134. Zhang, H.; Ryu, D.; Wu, Y.; Gariani, K.; Wang, X.; Luan, P.; D’Amico, D.; Ropelle, E.R.; Lutolf, M.P.; Aebersold, R.; et al. NAD+ Repletion Improves Mitochondrial and Stem Cell Function and Enhances Life Span in Mice. Science 2016, 352, 1436–1443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Cantó, C.; Houtkooper, R.H.; Pirinen, E.; Youn, D.Y.; Oosterveer, M.H.; Cen, Y.; Fernandez-Marcos, P.J.; Yamamoto, H.; Andreux, P.A.; Cettour-Rose, P.; et al. The NAD(+) Precursor Nicotinamide Riboside Enhances Oxidative Metabolism and Protects against High-Fat Diet-Induced Obesity. Cell Metab. 2012, 15, 838–847. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Nacarelli, T.; Lau, L.; Fukumoto, T.; Zundell, J.; Fatkhutdinov, N.; Wu, S.; Aird, K.M.; Iwasaki, O.; Kossenkov, A.V.; Schultz, D.; et al. NAD+ Metabolism Governs the Proinflammatory Senescence-Associated Secretome. Nat. Cell Biol. 2019, 21, 397–407. [Google Scholar] [CrossRef] [PubMed]
  137. Nassour, J.; Martien, S.; Martin, N.; Deruy, E.; Tomellini, E.; Malaquin, N.; Bouali, F.; Sabatier, L.; Wernert, N.; Pinte, S.; et al. Defective DNA Single-Strand Break Repair Is Responsible for Senescence and Neoplastic Escape of Epithelial Cells. Nat. Commun. 2016, 7, 10399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Ghosh, R.; Roy, S.; Franco, S. PARP1 Depletion Induces RIG-I-Dependent Signaling in Human Cancer Cells. PLoS ONE 2018, 13, e0194611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Wang, Z.; Gao, J.; Zhou, J.; Liu, H.; Xu, C. Olaparib Induced Senescence under P16 or P53 Dependent Manner in Ovarian Cancer. J. Gynecol. Oncol. 2018, 30, e26. [Google Scholar] [CrossRef]
  140. Ohanna, M.; Giuliano, S.; Bonet, C.; Imbert, V.; Hofman, V.; Zangari, J.; Bille, K.; Robert, C.; Bressac-de Paillerets, B.; Hofman, P.; et al. Senescent Cells Develop a PARP-1 and Nuclear Factor-{kappa}B-Associated Secretome (PNAS). Genes Dev. 2011, 25, 1245–1261. [Google Scholar] [CrossRef] [Green Version]
  141. Yu, W.; Dittenhafer-Reed, K.E.; Denu, J.M. SIRT3 Protein Deacetylates Isocitrate Dehydrogenase 2 (IDH2) and Regulates Mitochondrial Redox Status. J. Biol. Chem. 2012, 287, 14078–14086. [Google Scholar] [CrossRef] [Green Version]
  142. Igelmann, S.; Lessard, F.; Uchenunu, O.; Bouchard, J.; Fernandez-Ruiz, A.; Rowell, M.-C.; Lopes-Paciencia, S.; Papadopoli, D.; Fouillen, A.; Ponce, K.J.; et al. A Hydride Transfer Complex Reprograms NAD Metabolism and Bypasses Senescence. Mol. Cell 2021, 81, 3848–3865.e19. [Google Scholar] [CrossRef]
  143. Tarragó, M.G.; Chini, C.C.S.; Kanamori, K.S.; Warner, G.M.; Caride, A.; de Oliveira, G.C.; Rud, M.; Samani, A.; Hein, K.Z.; Huang, R.; et al. A Potent and Specific CD38 Inhibitor Ameliorates Age-Related Metabolic Dysfunction by Reversing Tissue NAD+ Decline. Cell Metab. 2018, 27, 1081–1095.e10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Zhang, D.; Hu, X.; Li, J.; Liu, J.; Baks-Te Bulte, L.; Wiersma, M.; Malik, N.-U.-A.; van Marion, D.M.S.; Tolouee, M.; Hoogstra-Berends, F.; et al. DNA Damage-Induced PARP1 Activation Confers Cardiomyocyte Dysfunction through NAD+ Depletion in Experimental Atrial Fibrillation. Nat. Commun. 2019, 10, 1307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Lin, K.Y.; Kraus, W.L. PARP Inhibitors for Cancer Therapy. Cell 2017, 169, 183. [Google Scholar] [CrossRef]
  146. Yap, T.A.; Plummer, R.; Azad, N.S.; Helleday, T. The DNA Damaging Revolution: PARP Inhibitors and Beyond. Am. Soc. Clin. Oncol. Educ. Book 2019, 39, 185–195. [Google Scholar] [CrossRef]
  147. Adashek, J.J.; Jain, R.K.; Zhang, J. Clinical Development of PARP Inhibitors in Treating Metastatic Castration-Resistant Prostate Cancer. Cells 2019, 8, 860. [Google Scholar] [CrossRef] [Green Version]
  148. Rajman, L.; Chwalek, K.; Sinclair, D.A. Therapeutic Potential of NAD-Boosting Molecules: The In Vivo Evidence. Cell Metab. 2018, 27, 529–547. [Google Scholar] [CrossRef] [Green Version]
  149. Migliavacca, E.; Tay, S.K.H.; Patel, H.P.; Sonntag, T.; Civiletto, G.; McFarlane, C.; Forrester, T.; Barton, S.J.; Leow, M.K.; Antoun, E.; et al. Mitochondrial Oxidative Capacity and NAD+ Biosynthesis Are Reduced in Human Sarcopenia across Ethnicities. Nat. Commun. 2019, 10, 5808. [Google Scholar] [CrossRef] [Green Version]
  150. Yoshino, J. ACMSD: A Novel Target for Modulating NAD+ Homeostasis. Trends Endocrinol. Metab. 2019, 30, 229–232. [Google Scholar] [CrossRef] [PubMed]
  151. Zhang, B.; Cui, S.; Bai, X.; Zhuo, L.; Sun, X.; Hong, Q.; Fu, B.; Wang, J.; Chen, X.; Cai, G. SIRT3 Overexpression Antagonizes High Glucose Accelerated Cellular Senescence in Human Diploid Fibroblasts via the SIRT3-FOXO1 Signaling Pathway. Age 2013, 35, 2237–2253. [Google Scholar] [CrossRef] [Green Version]
  152. Guo, Y.; Jia, X.; Cui, Y.; Song, Y.; Wang, S.; Geng, Y.; Li, R.; Gao, W.; Fu, D. Sirt3-Mediated Mitophagy Regulates AGEs-Induced BMSCs Senescence and Senile Osteoporosis. Redox Biol. 2021, 41, 101915. [Google Scholar] [CrossRef]
  153. Wątroba, M.; Szukiewicz, D. The Role of Sirtuins in Aging and Age-Related Diseases. Adv. Med. Sci. 2016, 61, 52–62. [Google Scholar] [CrossRef] [PubMed]
  154. Xia, W.; Geng, K. A Sirtuin Activator and an Anti-Inflammatory Molecule-Multifaceted Roles of Adjudin and Its Potential Applications for Aging-Related Diseases. Semin. Cell Dev. Biol. 2016, 59, 71–78. [Google Scholar] [CrossRef] [PubMed]
  155. Geng, K.; Fu, N.; Yang, X.; Xia, W. Adjudin Delays Cellular Senescence through Sirt3-mediated Attenuation of ROS Production. Int. J. Mol. Med. 2018, 42, 3522–3529. [Google Scholar] [CrossRef] [Green Version]
  156. Pozzan, T.; Rizzuto, R.; Volpe, P.; Meldolesi, J. Molecular and Cellular Physiology of Intracellular Calcium Stores. Physiol. Rev. 1994, 74, 595–636. [Google Scholar] [CrossRef]
  157. Rutter, G.A.; Rizzuto, R. Regulation of Mitochondrial Metabolism by ER Ca2+ Release: An Intimate Connection. Trends Biochem. Sci. 2000, 25, 215–221. [Google Scholar] [CrossRef]
  158. Rizzuto, R.; Bernardi, P.; Pozzan, T. Mitochondria as All-Round Players of the Calcium Game. J. Physiol. 2000, 529 Pt 1, 37–47. [Google Scholar] [CrossRef]
  159. Peng, T.-I.; Jou, M.-J. Oxidative Stress Caused by Mitochondrial Calcium Overload. Ann. N. Y. Acad. Sci. 2010, 1201, 183–188. [Google Scholar] [CrossRef] [PubMed]
  160. 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]
  161. Wong, R.; Steenbergen, C.; Murphy, E. Mitochondrial Permeability Transition Pore and Calcium Handling. Methods Mol. Biol. 2012, 810, 235–242. [Google Scholar] [CrossRef] [Green Version]
  162. Kwong, J.Q.; Molkentin, J.D. Physiological and Pathological Roles of the Mitochondrial Permeability Transition Pore in the Heart. Cell Metab. 2015, 21, 206–214. [Google Scholar] [CrossRef]
  163. 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] [PubMed]
  164. Scorrano, L.; De Matteis, M.A.; Emr, S.; Giordano, F.; Hajnóczky, G.; Kornmann, B.; Lackner, L.L.; Levine, T.P.; Pellegrini, L.; Reinisch, K.; et al. Coming Together to Define Membrane Contact Sites. Nat. Commun. 2019, 10, 1287. [Google Scholar] [CrossRef] [Green Version]
  165. 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] [PubMed] [Green Version]
  166. Ziegler, D.V.; Vindrieux, D.; Goehrig, D.; Jaber, S.; Collin, G.; Griveau, A.; Wiel, C.; Bendridi, N.; Djebali, S.; Farfariello, V.; et al. Calcium Channel ITPR2 and Mitochondria-ER Contacts Promote Cellular Senescence and Aging. Nat. Commun. 2021, 12, 720. [Google Scholar] [CrossRef] [PubMed]
  167. Farfariello, V.; Gordienko, D.V.; Mesilmany, L.; Touil, Y.; Germain, E.; Fliniaux, I.; Desruelles, E.; Gkika, D.; Roudbaraki, M.; Shapovalov, G.; et al. TRPC3 Shapes the ER-Mitochondria Ca2+ Transfer Characterizing Tumour-Promoting Senescence. Nat. Commun. 2022, 13, 956. [Google Scholar] [CrossRef]
  168. de J García-Rivas, G.; Carvajal, K.; Correa, F.; Zazueta, C. Ru360, a Specific Mitochondrial Calcium Uptake Inhibitor, Improves Cardiac Post-Ischaemic Functional Recovery in Rats in Vivo. Br. J. Pharmacol. 2006, 149, 829–837. [Google Scholar] [CrossRef] [Green Version]
  169. Novorolsky, R.J.; Nichols, M.; Kim, J.S.; Pavlov, E.V.; J Woods, J.; Wilson, J.J.; Robertson, G.S. The Cell-Permeable Mitochondrial Calcium Uniporter Inhibitor Ru265 Preserves Cortical Neuron Respiration after Lethal Oxygen Glucose Deprivation and Reduces Hypoxic/Ischemic Brain Injury. J. Cereb. Blood Flow Metab. 2020, 40, 1172–1181. [Google Scholar] [CrossRef]
  170. Increased Mitochondrial Ca2+ Contributes to Health Decline with Age and Duchene Muscular Dystrophy in C. elegans. BioRxiv. Available online: https://www.biorxiv.org/content/10.1101/2022.07.08.499319v1.full (accessed on 17 October 2022).
  171. Hoppel, C.L.; Lesnefsky, E.J.; Chen, Q.; Tandler, B. Mitochondrial Dysfunction in Cardiovascular Aging. Adv. Exp. Med. Biol. 2017, 982, 451–464. [Google Scholar] [CrossRef]
  172. Balaban, R.S.; Nemoto, S.; Finkel, T. Mitochondria, Oxidants, and Aging. Cell 2005, 120, 483–495. [Google Scholar] [CrossRef] [Green Version]
  173. Genova, M.L.; Lenaz, G. The Interplay Between Respiratory Supercomplexes and ROS in Aging. Antioxid. Redox Signal. 2015, 23, 208–238. [Google Scholar] [CrossRef]
  174. 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] [PubMed]
  175. Ahmed, E.K.; Rogowska-Wrzesinska, A.; Roepstorff, P.; Bulteau, A.-L.; Friguet, B. Protein Modification and Replicative Senescence of WI-38 Human Embryonic Fibroblasts. Aging Cell 2010, 9, 252–272. [Google Scholar] [CrossRef]
  176. Anisimov, V.N.; Bakeeva, L.E.; Egormin, P.A.; Filenko, O.F.; Isakova, E.F.; Manskikh, V.N.; Mikhelson, V.M.; Panteleeva, A.A.; Pasyukova, E.G.; Pilipenko, D.I.; et al. Mitochondria-Targeted Plastoquinone Derivatives as Tools to Interrupt Execution of the Aging Program. 5. SkQ1 Prolongs Lifespan and Prevents Development of Traits of Senescence. Biochemistry 2008, 73, 1329–1342. [Google Scholar] [CrossRef] [PubMed]
  177. Anisimov, V.N.; Egorov, M.V.; Krasilshchikova, M.S.; Lyamzaev, K.G.; Manskikh, V.N.; Moshkin, M.P.; Novikov, E.A.; Popovich, I.G.; Rogovin, K.A.; Shabalina, I.G.; et al. Effects of the Mitochondria-Targeted Antioxidant SkQ1 on Lifespan of Rodents. Aging 2011, 3, 1110–1119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Miura, S.; Saitoh, S.-I.; Kokubun, T.; Owada, T.; Yamauchi, H.; Machii, H.; Takeishi, Y. Mitochondrial-Targeted Antioxidant Maintains Blood Flow, Mitochondrial Function, and Redox Balance in Old Mice Following Prolonged Limb Ischemia. Int. J. Mol. Sci. 2017, 18, 1897. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Owada, T.; Yamauchi, H.; Saitoh, S.-I.; Miura, S.; Machii, H.; Takeishi, Y. Resolution of Mitochondrial Oxidant Stress Improves Aged-Cardiovascular Performance. Coron. Artery Dis. 2017, 28, 33–43. [Google Scholar] [CrossRef] [Green Version]
  180. Escames, G.; López, A.; García, J.A.; García, L.; Acuña-Castroviejo, D.; García, J.J.; López, L.C. The Role of Mitochondria in Brain Aging and the Effects of Melatonin. Curr. Neuropharmacol. 2010, 8, 182–193. [Google Scholar] [CrossRef] [Green Version]
  181. Migliaccio, E.; Giorgio, M.; Mele, S.; Pelicci, G.; Reboldi, P.; Pandolfi, P.P.; Lanfrancone, L.; Pelicci, P.G. The P66shc Adaptor Protein Controls Oxidative Stress Response and Life Span in Mammals. Nature 1999, 402, 309–313. [Google Scholar] [CrossRef]
  182. Napoli, C.; Martin-Padura, I.; de Nigris, F.; Giorgio, M.; Mansueto, G.; Somma, P.; Condorelli, M.; Sica, G.; De Rosa, G.; Pelicci, P. Deletion of the P66Shc Longevity Gene Reduces Systemic and Tissue Oxidative Stress, Vascular Cell Apoptosis, and Early Atherogenesis in Mice Fed a High-Fat Diet. Proc. Natl. Acad. Sci. USA 2003, 100, 2112–2116. [Google Scholar] [CrossRef] [Green Version]
  183. Pérez, V.I.; Van Remmen, H.; Bokov, A.; Epstein, C.J.; Vijg, J.; Richardson, A. The Overexpression of Major Antioxidant Enzymes Does Not Extend the Lifespan of Mice. Aging Cell 2009, 8, 73–75. [Google Scholar] [CrossRef]
  184. Van Remmen, H.; Ikeno, Y.; Hamilton, M.; Pahlavani, M.; Wolf, N.; Thorpe, S.R.; Alderson, N.L.; Baynes, J.W.; Epstein, C.J.; Huang, T.-T.; et al. Life-Long Reduction in MnSOD Activity Results in Increased DNA Damage and Higher Incidence of Cancer but Does Not Accelerate Aging. Physiol. Genom. 2003, 16, 29–37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Zhang, Y.; Ikeno, Y.; Qi, W.; Chaudhuri, A.; Li, Y.; Bokov, A.; Thorpe, S.R.; Baynes, J.W.; Epstein, C.; Richardson, A.; et al. Mice Deficient in Both Mn Superoxide Dismutase and Glutathione Peroxidase-1 Have Increased Oxidative Damage and a Greater Incidence of Pathology but No Reduction in Longevity. J. Gerontol. A Biol. Sci. Med. Sci. 2009, 64, 1212–1220. [Google Scholar] [CrossRef]
  186. Van Raamsdonk, J.M.; Hekimi, S. Deletion of the Mitochondrial Superoxide Dismutase Sod-2 Extends Lifespan in Caenorhabditis Elegans. PLoS Genet. 2009, 5, e1000361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Kornienko, J.S.; Smirnova, I.S.; Pugovkina, N.A.; Ivanova, J.S.; Shilina, M.A.; Grinchuk, T.M.; Shatrova, A.N.; Aksenov, N.D.; Zenin, V.V.; Nikolsky, N.N.; et al. High Doses of Synthetic Antioxidants Induce Premature Senescence in Cultivated Mesenchymal Stem Cells. Sci. Rep. 2019, 9, 1296. [Google Scholar] [CrossRef] [Green Version]
  188. Lambeth, J.D. NOX Enzymes and the Biology of Reactive Oxygen. Nat. Rev. Immunol. 2004, 4, 181–189. [Google Scholar] [CrossRef] [PubMed]
  189. Brand, M.D. The Sites and Topology of Mitochondrial Superoxide Production. Exp. Gerontol. 2010, 45, 466–472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Rhee, S.G. Cell Signaling. H2O2, a Necessary Evil for Cell Signaling. Science 2006, 312, 1882–1883. [Google Scholar] [CrossRef]
  191. Ristow, M.; Schmeisser, S. Extending Life Span by Increasing Oxidative Stress. Free Radic. Biol. Med. 2011, 51, 327–336. [Google Scholar] [CrossRef] [Green Version]
  192. Juntilla, M.M.; Patil, V.D.; Calamito, M.; Joshi, R.P.; Birnbaum, M.J.; Koretzky, G.A. AKT1 and AKT2 Maintain Hematopoietic Stem Cell Function by Regulating Reactive Oxygen Species. Blood 2010, 115, 4030–4038. [Google Scholar] [CrossRef] [Green Version]
  193. Owusu-Ansah, E.; Banerjee, U. Reactive Oxygen Species Prime Drosophila Haematopoietic Progenitors for Differentiation. Nature 2009, 461, 537–541. [Google Scholar] [CrossRef]
  194. Tormos, K.V.; Anso, E.; Hamanaka, R.B.; Eisenbart, J.; Joseph, J.; Kalyanaraman, B.; Chandel, N.S. Mitochondrial Complex III ROS Regulate Adipocyte Differentiation. Cell Metab. 2011, 14, 537–544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Malinska, D.; Kudin, A.P.; Bejtka, M.; Kunz, W.S. Changes in Mitochondrial Reactive Oxygen Species Synthesis during Differentiation of Skeletal Muscle Cells. Mitochondrion 2012, 12, 144–148. [Google Scholar] [CrossRef] [PubMed]
  196. Morimoto, H.; Iwata, K.; Ogonuki, N.; Inoue, K.; Atsuo, O.; Kanatsu-Shinohara, M.; Morimoto, T.; Yabe-Nishimura, C.; Shinohara, T. ROS Are Required for Mouse Spermatogonial Stem Cell Self-Renewal. Cell Stem Cell 2013, 12, 774–786. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Le Belle, J.E.; Orozco, N.M.; Paucar, A.A.; Saxe, J.P.; Mottahedeh, J.; Pyle, A.D.; Wu, H.; Kornblum, H.I. Proliferative Neural Stem Cells Have High Endogenous ROS Levels That Regulate Self-Renewal and Neurogenesis in a PI3K/Akt-Dependant Manner. Cell Stem Cell 2011, 8, 59–71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Tothova, Z.; Kollipara, R.; Huntly, B.J.; Lee, B.H.; Castrillon, D.H.; Cullen, D.E.; McDowell, E.P.; Lazo-Kallanian, S.; Williams, I.R.; Sears, C.; et al. FoxOs Are Critical Mediators of Hematopoietic Stem Cell Resistance to Physiologic Oxidative Stress. Cell 2007, 128, 325–339. [Google Scholar] [CrossRef] [Green Version]
  199. Ito, K.; Hirao, A.; Arai, F.; Matsuoka, S.; Takubo, K.; Hamaguchi, I.; Nomiyama, K.; Hosokawa, K.; Sakurada, K.; Nakagata, N.; et al. Regulation of Oxidative Stress by ATM Is Required for Self-Renewal of Haematopoietic Stem Cells. Nature 2004, 431, 997–1002. [Google Scholar] [CrossRef]
  200. Basso, E.; Fante, L.; Fowlkes, J.; Petronilli, V.; Forte, M.A.; Bernardi, P. Properties of the Permeability Transition Pore in Mitochondria Devoid of Cyclophilin D. J. Biol. Chem. 2005, 280, 18558–18561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Kinnally, K.W.; Peixoto, P.M.; Ryu, S.-Y.; Dejean, L.M. Is MPTP the Gatekeeper for Necrosis, Apoptosis, or Both? Biochim. Biophys. Acta 2011, 1813, 616–622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Baines, C.P.; Kaiser, R.A.; Purcell, N.H.; Blair, N.S.; Osinska, H.; Hambleton, M.A.; Brunskill, E.W.; Sayen, M.R.; Gottlieb, R.A.; Dorn, G.W.; et al. Loss of Cyclophilin D Reveals a Critical Role for Mitochondrial Permeability Transition in Cell Death. Nature 2005, 434, 658–662. [Google Scholar] [CrossRef]
  203. Ziegler, D.V.; Wiley, C.D.; Velarde, M.C. Mitochondrial Effectors of Cellular Senescence: Beyond the Free Radical Theory of Aging. Aging Cell 2015, 14, 1–7. [Google Scholar] [CrossRef]
  204. Bernardi, P.; Krauskopf, A.; Basso, E.; Petronilli, V.; Blachly-Dyson, E.; Blalchy-Dyson, E.; Di Lisa, F.; Forte, M.A. The Mitochondrial Permeability Transition from in Vitro Artifact to Disease Target. FEBS J. 2006, 273, 2077–2099. [Google Scholar] [CrossRef] [PubMed]
  205. Hafner, A.V.; Dai, J.; Gomes, A.P.; Xiao, C.-Y.; Palmeira, C.M.; Rosenzweig, A.; Sinclair, D.A. Regulation of the MPTP by SIRT3-Mediated Deacetylation of CypD at Lysine 166 Suppresses Age-Related Cardiac Hypertrophy. Aging 2010, 2, 914–923. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Picard, M.; Ritchie, D.; Thomas, M.M.; Wright, K.J.; Hepple, R.T. Alterations in Intrinsic Mitochondrial Function with Aging Are Fiber Type-Specific and Do Not Explain Differential Atrophy between Muscles. Aging Cell 2011, 10, 1047–1055. [Google Scholar] [CrossRef] [PubMed]
  207. Krestinina, O.; Azarashvili, T.; Baburina, Y.; Galvita, A.; Grachev, D.; Stricker, R.; Reiser, G. In Aging, the Vulnerability of Rat Brain Mitochondria Is Enhanced Due to Reduced Level of 2′,3′-Cyclic Nucleotide-3′-Phosphodiesterase (CNP) and Subsequently Increased Permeability Transition in Brain Mitochondria in Old Animals. Neurochem. Int. 2015, 80, 41–50. [Google Scholar] [CrossRef]
  208. Mather, M.; Rottenberg, H. Aging Enhances the Activation of the Permeability Transition Pore in Mitochondria. Biochem. Biophys. Res. Commun. 2000, 273, 603–608. [Google Scholar] [CrossRef] [Green Version]
  209. 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]
  210. Rottenberg, H.; Wu, S. Mitochondrial Dysfunction in Lymphocytes from Old Mice: Enhanced Activation of the Permeability Transition. Biochem. Biophys. Res. Commun. 1997, 240, 68–74. [Google Scholar] [CrossRef]
  211. Connern, C.P.; Halestrap, A.P. Recruitment of Mitochondrial Cyclophilin to the Mitochondrial Inner Membrane under Conditions of Oxidative Stress That Enhance the Opening of a Calcium-Sensitive Non-Specific Channel. Biochem. J. 1994, 302 (Pt 2), 321–324. [Google Scholar] [CrossRef] [Green Version]
  212. Martin, L.J.; Semenkow, S.; Hanaford, A.; Wong, M. Mitochondrial Permeability Transition Pore Regulates Parkinson’s Disease Development in Mutant α-Synuclein Transgenic Mice. Neurobiol. Aging 2014, 35, 1132–1152. [Google Scholar] [CrossRef] [Green Version]
  213. Martin, L.J.; Fancelli, D.; Wong, M.; Niedzwiecki, M.; Ballarini, M.; Plyte, S.; Chang, Q. GNX-4728, a Novel Small Molecule Drug Inhibitor of Mitochondrial Permeability Transition, Is Therapeutic in a Mouse Model of Amyotrophic Lateral Sclerosis. Front. Cell Neurosci. 2014, 8, 433. [Google Scholar] [CrossRef]
  214. Martin, L.J.; Gertz, B.; Pan, Y.; Price, A.C.; Molkentin, J.D.; Chang, Q. The Mitochondrial Permeability Transition Pore in Motor Neurons: Involvement in the Pathobiology of ALS Mice. Exp. Neurol. 2009, 218, 333–346. [Google Scholar] [CrossRef] [Green Version]
  215. Shum, L.C.; White, N.S.; Nadtochiy, S.M.; de Bentley, K.L.M.; Brookes, P.S.; Jonason, J.H.; Eliseev, R.A. Cyclophilin D Knock-Out Mice Show Enhanced Resistance to Osteoporosis and to Metabolic Changes Observed in Aging Bone. PLoS ONE 2016, 11, e0155709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Pastorino, J.G.; Tafani, M.; Rothman, R.J.; Marcinkeviciute, A.; Hoek, J.B.; Farber, J.L.; Marcineviciute, A. Functional Consequences of the Sustained or Transient Activation by Bax of the Mitochondrial Permeability Transition Pore. J. Biol. Chem. 1999, 274, 31734–31739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Elmore, S. Apoptosis: A Review of Programmed Cell Death. Toxicol. Pathol. 2007, 35, 495–516. [Google Scholar] [CrossRef] [PubMed]
  218. Salvador-Gallego, R.; Mund, M.; Cosentino, K.; Schneider, J.; Unsay, J.; Schraermeyer, U.; Engelhardt, J.; Ries, J.; García-Sáez, A.J. Bax Assembly into Rings and Arcs in Apoptotic Mitochondria Is Linked to Membrane Pores. EMBO J. 2016, 35, 389–401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Große, L.; Wurm, C.A.; Brüser, C.; Neumann, D.; Jans, D.C.; Jakobs, S. Bax Assembles into Large Ring-like Structures Remodeling the Mitochondrial Outer Membrane in Apoptosis. EMBO J. 2016, 35, 402–413. [Google Scholar] [CrossRef] [PubMed]
  220. Tait, S.W.G.; Green, D.R. Mitochondrial Regulation of Cell Death. Cold Spring Harb. Perspect. Biol. 2013, 5, a008706. [Google Scholar] [CrossRef] [Green Version]
  221. Chinnaiyan, A.M. The Apoptosome: Heart and Soul of the Cell Death Machine. Neoplasia 1999, 1, 5–15. [Google Scholar] [CrossRef] [Green Version]
  222. Hill, M.M.; Adrain, C.; Duriez, P.J.; Creagh, E.M.; Martin, S.J. Analysis of the Composition, Assembly Kinetics and Activity of Native Apaf-1 Apoptosomes. EMBO J. 2004, 23, 2134–2145. [Google Scholar] [CrossRef] [Green Version]
  223. Borras, C.; M Abdelaziz, K.; Gambini, J.; Serna, E.; Inglés, M.; de la Fuente, M.; Garcia, I.; Matheu, A.; Sanchís, P.; Belenguer, A.; et al. Human Exceptional Longevity: Transcriptome from Centenarians Is Distinct from Septuagenarians and Reveals a Role of Bcl-XL in Successful Aging. Aging 2016, 8, 3185–3201. [Google Scholar] [CrossRef]
  224. Borrás, C.; Mas-Bargues, C.; Román-Domínguez, A.; Sanz-Ros, J.; Gimeno-Mallench, L.; Inglés, M.; Gambini, J.; Viña, J. BCL-XL, a Mitochondrial Protein Involved in Successful Aging: From C. Elegans to Human Centenarians. Int. J. Mol. Sci. 2020, 21, 418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Rochette, P.J.; Brash, D.E. Progressive Apoptosis Resistance Prior to Senescence and Control by the Anti-Apoptotic Protein BCL-XL. Mech. Ageing Dev. 2008, 129, 207–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Zhu, Y.; Tchkonia, T.; Fuhrmann-Stroissnigg, H.; Dai, H.M.; Ling, Y.Y.; Stout, M.B.; Pirtskhalava, T.; Giorgadze, N.; Johnson, K.O.; Giles, C.B.; et al. Identification of a Novel Senolytic Agent, Navitoclax, Targeting the Bcl-2 Family of Anti-apoptotic Factors. Aging Cell 2016, 15, 428–435. [Google Scholar] [CrossRef]
  227. Chang, J.; Wang, Y.; Shao, L.; Laberge, R.-M.; Demaria, M.; Campisi, J.; Janakiraman, K.; Sharpless, N.E.; Ding, S.; Feng, W.; et al. Clearance of Senescent Cells by ABT263 Rejuvenates Aged Hematopoietic Stem Cells in Mice. Nat. Med. 2016, 22, 78–83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Pemmaraju, N.; Garcia, J.S.; Potluri, J.; Harb, J.G.; Sun, Y.; Jung, P.; Qin, Q.Q.; Tantravahi, S.K.; Verstovsek, S.; Harrison, C. Addition of Navitoclax to Ongoing Ruxolitinib Treatment in Patients with Myelofibrosis (REFINE): A Post-Hoc Analysis of Molecular Biomarkers in a Phase 2 Study. Lancet Haematol. 2022, 9, e434–e444. [Google Scholar] [CrossRef] [PubMed]
  229. Harrison, C.N.; Garcia, J.S.; Somervaille, T.C.P.; Foran, J.M.; Verstovsek, S.; Jamieson, C.; Mesa, R.; Ritchie, E.K.; Tantravahi, S.K.; Vachhani, P.; et al. Addition of Navitoclax to Ongoing Ruxolitinib Therapy for Patients With Myelofibrosis With Progression or Suboptimal Response: Phase II Safety and Efficacy. J. Clin. Oncol. 2022, 40, 1671–1680. [Google Scholar] [CrossRef]
  230. Goldberg, A.D.; Allis, C.D.; Bernstein, E. Epigenetics: A Landscape Takes Shape. Cell 2007, 128, 635–638. [Google Scholar] [CrossRef] [Green Version]
  231. Muñoz-Najar, U.; Sedivy, J.M. Epigenetic Control of Aging. Antioxid. Redox Signal. 2011, 14, 241–259. [Google Scholar] [CrossRef] [Green Version]
  232. Bollati, V.; Schwartz, J.; Wright, R.; Litonjua, A.; Tarantini, L.; Suh, H.; Sparrow, D.; Vokonas, P.; Baccarelli, A. Decline in Genomic DNA Methylation through Aging in a Cohort of Elderly Subjects. Mech. Ageing Dev. 2009, 130, 234–239. [Google Scholar] [CrossRef] [Green Version]
  233. Bjornsson, H.T.; Sigurdsson, M.I.; Fallin, M.D.; Irizarry, R.A.; Aspelund, T.; Cui, H.; Yu, W.; Rongione, M.A.; Ekström, T.J.; Harris, T.B.; et al. Intra-Individual Change over Time in DNA Methylation with Familial Clustering. JAMA 2008, 299, 2877–2883. [Google Scholar] [CrossRef]
  234. Horvath, S. DNA Methylation Age of Human Tissues and Cell Types. Genome. Biol. 2013, 14, R115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Szafranski, K.; Abraham, K.J.; Mekhail, K. Non-Coding RNA in Neural Function, Disease, and Aging. Front. Genet. 2015, 6, 87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Inukai, S.; de Lencastre, A.; Turner, M.; Slack, F. Novel MicroRNAs Differentially Expressed during Aging in the Mouse Brain. PLoS ONE 2012, 7, e40028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Persengiev, S.; Kondova, I.; Otting, N.; Koeppen, A.H.; Bontrop, R.E. Genome-Wide Analysis of MiRNA Expression Reveals a Potential Role for MiR-144 in Brain Aging and Spinocerebellar Ataxia Pathogenesis. Neurobiol. Aging 2011, 32, 2316.e17–2316.e27. [Google Scholar] [CrossRef] [PubMed]
  238. Narita, M.; Nũnez, S.; Heard, E.; Narita, M.; Lin, A.W.; Hearn, S.A.; Spector, D.L.; Hannon, G.J.; Lowe, S.W. Rb-Mediated Heterochromatin Formation and Silencing of E2F Target Genes during Cellular Senescence. Cell 2003, 113, 703–716. [Google Scholar] [CrossRef] [Green Version]
  239. O’Sullivan, R.J.; Kubicek, S.; Schreiber, S.L.; Karlseder, J. Reduced Histone Biosynthesis and Chromatin Changes Arising from a Damage Signal at Telomeres. Nat. Struct. Mol. Biol. 2010, 17, 1218–1225. [Google Scholar] [CrossRef] [Green Version]
  240. Giorgio, E.D.; Paluvai, H.; Dalla, E.; Ranzino, L.; Renzini, A.; Moresi, V.; Minisini, M.; Picco, R.; Brancolini, C. HDAC4 Degradation during Senescence Unleashes an Epigenetic Program Driven by AP-1/P300 at Selected Enhancers and Super-Enhancers. Genome Biol. 2021, 22, 129. [Google Scholar] [CrossRef]
  241. Capell, B.C.; Drake, A.M.; Zhu, J.; Shah, P.P.; Dou, Z.; Dorsey, J.; Simola, D.F.; Donahue, G.; Sammons, M.; Rai, T.S.; et al. MLL1 Is Essential for the Senescence-Associated Secretory Phenotype. Genes Dev. 2016, 30, 321–336. [Google Scholar] [CrossRef] [Green Version]
  242. Shaughnessy, D.T.; McAllister, K.; Worth, L.; Haugen, A.C.; Meyer, J.N.; Domann, F.E.; Van Houten, B.; Mostoslavsky, R.; Bultman, S.J.; Baccarelli, A.A.; et al. Mitochondria, Energetics, Epigenetics, and Cellular Responses to Stress. Environ. Health Perspect 2014, 122, 1271–1278. [Google Scholar] [CrossRef]
  243. Martínez-Reyes, I.; Chandel, N.S. Mitochondrial TCA Cycle Metabolites Control Physiology and Disease. Nat. Commun. 2020, 11, 102. [Google Scholar] [CrossRef]
  244. Salminen, A.; Kaarniranta, K.; Hiltunen, M.; Kauppinen, A. Krebs Cycle Dysfunction Shapes Epigenetic Landscape of Chromatin: Novel Insights into Mitochondrial Regulation of Aging Process. Cell Signal. 2014, 26, 1598–1603. [Google Scholar] [CrossRef]
  245. Schroeder, E.A.; Raimundo, N.; Shadel, G.S. Epigenetic Silencing Mediates Mitochondria Stress-Induced Longevity. Cell Metab. 2013, 17, 954–964. [Google Scholar] [CrossRef] [Green Version]
  246. Hayakawa, K.; Esposito, E.; Wang, X.; Terasaki, Y.; Liu, Y.; Xing, C.; Ji, X.; Lo, E.H. Transfer of Mitochondria from Astrocytes to Neurons after Stroke. Nature 2016, 535, 551–555. [Google Scholar] [CrossRef] [Green Version]
  247. Gurke, S.; Barroso, J.F.V.; Gerdes, H.-H. The Art of Cellular Communication: Tunneling Nanotubes Bridge the Divide. Histochem. Cell Biol. 2008, 129, 539–550. [Google Scholar] [CrossRef] [Green Version]
  248. Saha, T.; Dash, C.; Jayabalan, R.; Khiste, S.; Kulkarni, A.; Kurmi, K.; Mondal, J.; Majumder, P.K.; Bardia, A.; Jang, H.L.; et al. Intercellular Nanotubes Mediate Mitochondrial Trafficking between Cancer and Immune Cells. Nat. Nanotechnol. 2022, 17, 98–106. [Google Scholar] [CrossRef]
  249. Islam, M.N.; Das, S.R.; Emin, M.T.; Wei, M.; Sun, L.; Westphalen, K.; Rowlands, D.J.; Quadri, S.K.; Bhattacharya, S.; Bhattacharya, J. Mitochondrial Transfer from Bone-Marrow-Derived Stromal Cells to Pulmonary Alveoli Protects against Acute Lung Injury. Nat. Med. 2012, 18, 759–765. [Google Scholar] [CrossRef] [Green Version]
  250. Lightowlers, R.N.; Chrzanowska-Lightowlers, Z.M.; Russell, O.M. Mitochondrial Transplantation—A Possible Therapeutic for Mitochondrial Dysfunction? EMBO Rep. 2020, 21, e50964. [Google Scholar] [CrossRef]
  251. Kaza, A.K.; Wamala, I.; Friehs, I.; Kuebler, J.D.; Rathod, R.H.; Berra, I.; Ericsson, M.; Yao, R.; Thedsanamoorthy, J.K.; Zurakowski, D.; et al. Myocardial Rescue with Autologous Mitochondrial Transplantation in a Porcine Model of Ischemia/Reperfusion. J. Thorac. Cardiovasc. Surg. 2017, 153, 934–943. [Google Scholar] [CrossRef] [Green Version]
  252. Doulamis, I.P.; Guariento, A.; Duignan, T.; Kido, T.; Orfany, A.; Saeed, M.Y.; Weixler, V.H.; Blitzer, D.; Shin, B.; Snay, E.R.; et al. Mitochondrial Transplantation by Intra-Arterial Injection for Acute Kidney Injury. Am. J. Physiol. Renal. Physiol. 2020, 319, F403–F413. [Google Scholar] [CrossRef]
  253. Moskowitzova, K.; Orfany, A.; Liu, K.; Ramirez-Barbieri, G.; Thedsanamoorthy, J.K.; Yao, R.; Guariento, A.; Doulamis, I.P.; Blitzer, D.; Shin, B.; et al. Mitochondrial Transplantation Enhances Murine Lung Viability and Recovery after Ischemia-Reperfusion Injury. Am. J. Physiol. Lung Cell Mol. Physiol. 2020, 318, L78–L88. [Google Scholar] [CrossRef]
  254. Adlimoghaddam, A.; Benson, T.; Albensi, B.C. Mitochondrial Transfusion Improves Mitochondrial Function Through Up-Regulation of Mitochondrial Complex II Protein Subunit SDHB in the Hippocampus of Aged Mice. Mol. Neurobiol. 2022, 59, 6009–6017. [Google Scholar] [CrossRef] [PubMed]
  255. Ali Pour, P.; Kenney, M.C.; Kheradvar, A. Bioenergetics Consequences of Mitochondrial Transplantation in Cardiomyocytes. J. Am. Heart Assoc. 2020, 9, e014501. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Cellular senescence is a terminal state of proliferation arrest in response to stressors or damages. Classic markers of cellular senescence are increased cell size, development of a senescence-associated secretory phenotype (SASP), chromatin remodelling, increase (↑) in the lysosomal compartment and autophagy, and mitochondrial alterations. Created with BioRender.com.
Figure 1. Cellular senescence is a terminal state of proliferation arrest in response to stressors or damages. Classic markers of cellular senescence are increased cell size, development of a senescence-associated secretory phenotype (SASP), chromatin remodelling, increase (↑) in the lysosomal compartment and autophagy, and mitochondrial alterations. Created with BioRender.com.
Pharmaceutics 15 00352 g001
Figure 2. Mitochondrial dysfunctions are a hallmark of both ageing and cellular senescence and represent crucial targets in treating ageing-associated diseases. (A) Both aged and senescent cells are characterised by modifications in the mitochondrial network, dynamics, and interactions between the organelle and the rest of the cell. However, differences can be observed between the two conditions. In particular, modifications in the mitochondria-to-nucleus retrograde responses can induce different epigenetic changes in senescent versus old cells and modulate signal transcription pathways, such as the mitochondrial unfolded protein response (UPRmt), differently. Mitochondrial DNA (mtDNA) release, moreover, can be detected and cause inflammation in both conditions, but its role is particularly relevant as part of the senescence-associated secretory phenotype (SASP). (B) Most mitochondrial functions are similarly defective or altered in aged and senescent cells. Inefficient mitochondrial respiration results in impaired metabolism, a drop in NAD+ levels, and ROS production. Calcium and ROS accumulation might cause increased susceptibility to the mitochondria permeability transition pore (mPTP) opening and apoptosis. Arrows indicate increase (↑) and decrease (↓) of the correspondent feature. Created with BioRender.com.
Figure 2. Mitochondrial dysfunctions are a hallmark of both ageing and cellular senescence and represent crucial targets in treating ageing-associated diseases. (A) Both aged and senescent cells are characterised by modifications in the mitochondrial network, dynamics, and interactions between the organelle and the rest of the cell. However, differences can be observed between the two conditions. In particular, modifications in the mitochondria-to-nucleus retrograde responses can induce different epigenetic changes in senescent versus old cells and modulate signal transcription pathways, such as the mitochondrial unfolded protein response (UPRmt), differently. Mitochondrial DNA (mtDNA) release, moreover, can be detected and cause inflammation in both conditions, but its role is particularly relevant as part of the senescence-associated secretory phenotype (SASP). (B) Most mitochondrial functions are similarly defective or altered in aged and senescent cells. Inefficient mitochondrial respiration results in impaired metabolism, a drop in NAD+ levels, and ROS production. Calcium and ROS accumulation might cause increased susceptibility to the mitochondria permeability transition pore (mPTP) opening and apoptosis. Arrows indicate increase (↑) and decrease (↓) of the correspondent feature. Created with BioRender.com.
Pharmaceutics 15 00352 g002
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

Protasoni, M.; Serrano, M. Targeting Mitochondria to Control Ageing and Senescence. Pharmaceutics 2023, 15, 352. https://doi.org/10.3390/pharmaceutics15020352

AMA Style

Protasoni M, Serrano M. Targeting Mitochondria to Control Ageing and Senescence. Pharmaceutics. 2023; 15(2):352. https://doi.org/10.3390/pharmaceutics15020352

Chicago/Turabian Style

Protasoni, Margherita, and Manuel Serrano. 2023. "Targeting Mitochondria to Control Ageing and Senescence" Pharmaceutics 15, no. 2: 352. https://doi.org/10.3390/pharmaceutics15020352

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