Next Article in Journal
Proteomic Analysis of the Meniscus Cartilage in Osteoarthritis
Next Article in Special Issue
Conserved and Distinct Elements of Phagocytosis in Human and C. elegans
Previous Article in Journal
Mitochondrial Modulations, Autophagy Pathways Shifts in Viral Infections: Consequences of COVID-19
Previous Article in Special Issue
Rapamycin Plus Doxycycline Combination Affects Growth Arrest and Selective Autophagy-Dependent Cell Death in Breast Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mitochondrial Dynamics and Mitophagy in Skeletal Muscle Health and Aging

by
Jean-Philippe Leduc-Gaudet
1,2,3,*,†,
Sabah N. A. Hussain
1,2,†,
Esther Barreiro
4,5,6,† and
Gilles Gouspillou
1,2,3,7,†
1
Research Institute of the McGill University Health Centre, Department of Critical Care, Montréal, QC H4A 3J1, Canada
2
Meakins-Christie Laboratories, Department of Medicine, McGill University, Montréal, QC H4A 3J1, Canada
3
Département des Sciences de l’activité Physique, Faculté des Sciences, UQAM, Montréal, QC H2X 1Y4, Canada
4
Pulmonology Department-Muscle Wasting & Cachexia in Chronic Respiratory Diseases & Lung Cancer Research Group, IMIM-Hospital del Mar, Parc de Salut Mar, Biomedical Research Park (PRBB), C/Dr. Aiguader, 88, 08003 Barcelona, Spain
5
Centro de Investigación en Red de Enfermedades Respiratorias (CIBERES), Instituto de Salud Carlos III (ISCIII), 28029 Madrid, Spain
6
Health and Experimental Sciences Department (CEXS), Pompeu Fabra University (UPF), Biomedical Research Park (PRBB), C/Dr. Aiguader, 88, 08003 Barcelona, Spain
7
Centre de Recherche de l’Institut Universitaire de Gériatrie de Montréal, Montréal, QC H3W 1W5, Canada
*
Author to whom correspondence should be addressed.
All authors contributed equally to this review.
Int. J. Mol. Sci. 2021, 22(15), 8179; https://doi.org/10.3390/ijms22158179
Submission received: 7 July 2021 / Revised: 23 July 2021 / Accepted: 27 July 2021 / Published: 30 July 2021
(This article belongs to the Special Issue Molecular Research of Autophagy and Apoptosis)

Abstract

:
The maintenance of mitochondrial integrity is critical for muscle health. Mitochondria, indeed, play vital roles in a wide range of cellular processes, including energy supply, Ca2+ homeostasis, retrograde signaling, cell death, and many others. All mitochondria-containing cells, including skeletal muscle cells, dispose of several pathways to maintain mitochondrial health, including mitochondrial biogenesis, mitochondrial-derived vesicles, mitochondrial dynamics (fusion and fission process shaping mitochondrial morphology), and mitophagy—the process in charge of the removal of mitochondria though autophagy. The loss of skeletal muscle mass (atrophy) is a major health problem worldwide, especially in older people. Currently, there is no treatment to counteract the progressive decline in skeletal muscle mass and strength that occurs with aging, a process termed sarcopenia. There is increasing data, including our own, suggesting that accumulation of dysfunctional mitochondria contributes to the development of sarcopenia. Impairments in mitochondrial dynamics and mitophagy were recently proposed to contribute to sarcopenia. This review summarizes the current state of knowledge on the role played by mitochondrial dynamics and mitophagy in skeletal muscle health and in the development of sarcopenia. We also highlight recent studies showing that enhancing mitophagy in skeletal muscle is a promising therapeutic target to prevent or even treat skeletal muscle dysfunction in the elderly.

1. Introduction

Skeletal muscle, which accounts for over 40% of total body mass, is highly adaptable and of critical importance for general health. Indeed, it plays key roles in posture, mobility, thermogenesis, glucose homeostasis, and energy metabolism. The regulation of skeletal muscle mass is mainly controlled by a fine balance between protein synthesis and protein degradation [1]. Skeletal muscle has a striking ability to adapt to various physiological and pathological conditions by changing its size, composition, and metabolic properties. Muscle atrophy, or the loss of muscle mass, is an important clinical problem that leads to negative health outcomes, especially during aging.
The progressive decline in muscle mass and strength that occurs with aging, a process known as sarcopenia, can have serious consequences on the lifestyle and functional status of the elderly. For instance, lower muscle mass and strength has been associated with functional impairment, falls, and physical disability [2,3]. Sarcopenia is a strong predictor of hospitalization and is highly linked to short- and long-term mortality in older adults [4,5,6]. Given the increasing population of older adults all over the world, it is not surprising that the age-related decline in muscle mass, strength, and function has become a major global health concern. Highlighting its serious health consequences, sarcopenia was recognized in 2016 as a muscle condition/disease and received its own ICD (international classification of diseases)-10-Clinical Modification code (M62.84) [7].
Several biological mechanisms have been proposed to explain the development of sarcopenia [8]. Among these mechanisms, strong experimental evidence indicates that the accumulation of mitochondrial dysfunction plays an important role in the muscle aging process. Indeed, many studies conducted in both rodents and humans have shown that aging is associated with a decline in mitochondrial bioenergetics [9,10,11]. Several studies have also shown that the mitochondrial permeability transition pore (mPTP) becomes dysfunctional in old rodents and humans [9,12,13,14]. Alterations in mPTP function are undoubtedly contributing to the aging-related loss of muscle mass and function, since mPTP opening can trigger several atrophy-regulating pathways. First, mPTP opening can result in mitochondrial reactive oxygen species (ROS) overproduction [15]. ROS can, in turn, trigger the muscle atrophy program by activating the Fork head box O (FoxO) transcription factor family [16,17]. Second, mPTP opening can result in mitochondrial DNA release and consequential activation of the NLRP3 (NOD-, LRR- and pyrin domain-containing protein 3) inflammasome, which was shown to increase the expression of ubiquitin ligases involved in muscle atrophy [18,19]. Third, cytochrome C release secondary to mPTP opening was shown to increase proteasomal activity secondary to caspase 3 activation [20]. Fourth, mPTP opening can lead to the release of pro-apoptotic factors such as endonuclease G, which was shown to accumulate in myonuclei of old individuals [9]. Further strengthening the view that mitochondria play key roles in the aging-related loss of muscle mass and function, the most efficient interventions to attenuate sarcopenia are exercise training [21,22,23,24] and calorie restriction [25,26,27], both well-known to improve mitochondrial health [28].
While the accumulation of mitochondrial dysfunction is now generally seen as a hallmark of muscle aging, the mechanisms underlying the progressive decline in mitochondrial function remain poorly understood. In this article, we critically review the available literature that has positioned impairments in mitochondrial dynamics and mitophagy as potential contributors. We also review recent evidence showing that targeting mitophagy might be a promising strategy to prevent or even treat sarcopenia.

2. The Importance of Mitochondria Dynamics for Skeletal Muscle Health

Initially considered as bean-shaped organelles, mitochondria are now known to display a complex architecture, especially in skeletal muscle cells [29]. Mitochondria are dynamic organelles that are constantly undergoing fission and fusion so as to adapt their morphology to the cellular environment (Figure 1). Mitochondrial fusion is mainly regulated by the Mitofusin 1 & 2 (Mfn 1 & 2) and OPtic Atrophy 1 (Opa 1) proteins. Mitochondrial fission is mainly governed by Dynamin-Related Protein 1 (Drp1) (reviewed in [30,31]). Several other proteins targeting the outer membrane, such as mitochondrial fission 1 protein (Fis1), mitochondrial fission factor (Mff), and mitochondrial dynamic protein 49 and 51 (MiD49, and MiD51), participate in the fragmentation of mitochondria [32]. Figure 1 shows the known regulators involved in mitochondrial fusion and fission.
In the past decade, several studies have established the importance of mitochondrial dynamics in maintaining skeletal muscle and mitochondrial integrity and function. It was, for instance, reported that mice lacking both Mfn-1 and Mfn-2 in skeletal muscle display severe mitochondrial dysfunction, accumulation of mitochondrial DNA damage, and a severe deficit in growth [33]. Mfn2 deletion in skeletal muscle was shown to result in oxidative stress and muscle atrophy in adult mice [34]. Deletion of both Mfn1 and Mfn2 in adult skeletal muscle resulted in a major decrease in exercise performance, indicating that mitochondrial fusion is essential to muscle performance in mice [35].
Specific deletion of Opa1 in skeletal muscle leads to mitochondrial dysfunction, oxidative stress, endoplasmic reticulum (ER) stress, and inflammation [36,37,38]. Opa1 deficiency also promotes the secretion of Fibroblast Growth Factor 21 (FGF21) from skeletal muscle, which leads to altered lipid homeostasis, inflammation, and the senescence of different tissues [36]. Taken altogether, the available literature clearly indicates that mitochondrial fusion proteins are essential for the maintenance of mitochondrial and skeletal muscle health.
Several recent studies have highlighted the physiological importance of mitochondrial fission in the maintenance of skeletal muscle health. Muscle-specific Drp1 overexpression was shown to impair skeletal muscle growth in mice [39,40]. Inhibition of mitochondrial fission by genetic silencing of Fis1 and Drp1 in skeletal muscle has been shown to protect against muscle wasting induced by starvation or by overexpression of the atrophying transcription factor FoxO3a [41]. Muscle-specific Drp1 knockout in mice leads to a severe myopathic phenotype including muscle wasting, weakness, and signs of muscle degeneration and regeneration [42]. The myopathic phenotype was evident about 2 months after muscle deletion of Drp1 and was also present at 6 months [42]. The importance of Drp1 in regulating muscle function was further emphasized by our group [43], which revealed that muscle-selective Drp1 knockdown decreases mitochondrial respiration and increases markers of muscle regeneration, denervation, fibrosis, and oxidative stress. Interestingly, both Drp1 knockout and knockdown altered autophagy and mitophagy levels [42,43]. As mentioned above, overexpression of Drp1 impaired skeletal muscle growth [39], whereas deletion or insufficient expression of Drp1 in skeletal muscle led to 40–50% atrophy [42,43], indicating that Drp1 is a key player in regulating skeletal muscle development and maintenance. Evidence from Drosophila also indicates that manipulating mitochondrial dynamics can impact myofibril development [44]. Another recent study showed that loss of Fis1 leads to mitochondrial dysfunction, proteostasis impairment, muscle degeneration, and reduced lifespan in Fis1 mutant flies [45]. Taken altogether, the findings discussed above clearly indicate that mitochondrial fission is essential for skeletal muscle health.

3. Mitochondrial Dynamics in the Skeletal Muscle Aging Process: Current State of Knowledge and Considerations for Future Research

As highlighted above, the available literature clearly positions mitochondrial dynamics as key processes in the maintenance of muscle and mitochondrial health. It is, therefore, not surprising that many studies have investigated whether impairment in mitochondrial dynamics could contribute to sarcopenia. However, and as will be reviewed below, the involvement of mitochondrial dynamics in the skeletal muscle aging process remains unclear as conflicting data exist in the field (Table 1).
In their elegant study, Sebastian et al. reported that mice lacking Mfn2 in skeletal muscle from birth display accelerated sarcopenia [34]. They also reported that aged control (wild type, WT) mice displayed a reduction in Mfn2 content [34]. However, the finding that Mfn2 content declines with aging is far from being universal. Indeed, Mfn2 content has been reported as decreased [34,49,55], unchanged [52], or even increased in aged rodent skeletal muscles [47,50,54]. In humans, while Mfn2 expression was decreased in aged skeletal muscle in one study [60], others have reported either similar [56,58,59,61] or even higher Mfn2 content in aged skeletal muscles [61]. Smartly conflicting data also exist for Drp1 and Opa1.
In a recent study, Tezze et al. reported that Opa1 content normalized to complex II was lower in muscle from old sedentary individuals but not in old sportsmen. Interestingly, Opa1 content normalized to complex II content correlated with muscle fiber diameter and specific force in elderly individuals [36]. Somewhat in line with this finding, Joseph et al. reported that Opa1 content was lower in the skeletal muscle of both low functioning and high functioning individuals [59]. However, several studies did not detect any impact of aging on Opa1 content in the skeletal muscle of aged mice [52,54], rats [50], and humans [57,58]. It is worth mentioning that Tezze et al. also reported that 18-month-old mice display lower levels of Opa1, but also showed that one week of exercise training was sufficient to restore Opa1 in aged skeletal muscles [36].
Conflicting data also surround the impact of aging on Drp1 content. Indeed, studies in rodents have reported increased [49,50,51] or unchanged [34,47,52,54] Drp1 content in aged skeletal muscles. The literature in humans also contains conflicting data, with some studies reporting a decreased [58] or an unchanged [59] Drp1 content in skeletal muscle of old individuals.
Clearly, no consensus on the role that mitochondrial dynamics play in the muscle aging process has been reached. Recently developed transgenic mouse lines with photoactivatable mitochondria [62] might prove useful in addressing this issue. Defining whether impairment in mitochondrial dynamics is a likely contributor to the aging-related loss of muscle mass and strength in humans will require longitudinal studies. Such studies will need to pay particular attention to physical activity levels of participants, as the latter can impact the expression of proteins regulating mitochondrial dynamics [63,64,65]. Further studies will also need to take into account the possibility that the impact of aging on mitochondrial dynamics and morphology might be fiber type-specific. Indeed, we recently reported in aged rats that the oxidative soleus muscle displayed signs of mitochondrial fragmentation, while the glycolytic white gastrocnemius displayed signs of mitochondrial elongation with aging [50]. Other studies have also highlighted that mitochondrial morphology and dynamics differ across muscle fiber types [61,66,67]. Importantly, a major gap in knowledge in the field that needs to be addressed urgently is whether sex differences exist in the role that mitochondrial dynamics play in the skeletal muscle aging process. Indeed, and to the best of our knowledge, no previous study has investigated whether the content of proteins regulating mitochondrial dynamics is differently affected in aged males vs. females.
Studies aiming at defining whether mitochondrial dynamics can be targeted to counteract sarcopenia are also needed. To date, the modulation (overexpression or deletion) of proteins regulating mitochondrial dynamics in mice has either shown no benefit to muscle cells [68] or was proven deleterious to muscle health (Drp1 overexpression, Drp1 knockout, Drp1 knockdown, and Mfn1+Mfn2 knockout). Strengthening the need for such mechanistic studies, promoting Drp1-mediated mitochondrial fission in middle-aged flies was shown to extend lifespan and improve muscle mitochondrial function, likely by facilitating mitophagy [69] (mitochondrial fission can, indeed, facilitate the removal of damaged/dysfunctional mitochondria through mitophagy [70]). This study, therefore, suggested that Drp1 overexpression could be a potential strategy to attenuate the aging-related accumulation of mitochondrial dysfunction and counter sarcopenia. However, we recently reported that knocking down or overexpressing Drp1 late in life in mouse skeletal muscle (from 18 to 22 months of age) is detrimental to skeletal muscle mass and mitochondrial health [71]. These recent findings, therefore, indicate that targeting Drp1 expression is unlikely to be a viable target to counter sarcopenia and indicate that mitochondrial fission must be maintained within a specific physiological range to preserve muscle and mitochondrial health. Whether enhancing/overexpressing mitochondrial fusion proteins, such as Mfn2 in muscles of aged WT mice, can attenuate the progression of sarcopenia has never been tested.

4. The Role of Autophagy in Skeletal Muscle Health and Aging

The word autophagy derives from two Greek words “auto”, meaning self, and “phagy”, meaning eating. Three different delivery systems have been described in mammals: macro-autophagy, chaperone-mediated autophagy, and micro-autophagy. In this review, the term autophagy is synonymous with macroautophagy, which is characterized by the formation of double-membrane vesicles known as autophagosomes that surround portions of cytoplasm, organelles, and protein aggregates that are subsequently delivered to lysosomes for degradation. The proper maintenance of this catabolic process is essential for survival and has important roles in various physiological functions [72,73]. Under normal conditions, autophagy mainly prevents the accumulation of damaged organelles and misfolded proteins. In response to stress, such as starvation, autophagy acts mainly as a pro-survival mechanism, by providing metabolic substrates such as amino acids through the bulk degradation of cytoplasmic components; however, excessive self-degradation can lead to cell death and pathological phenotypes [74,75].
The formation of autophagosomes and the degradation of its cargo involve several steps including initiation, elongation, maturation, fusion, and degradation (Figure 2). These steps require energy and involve more than 30 autophagy-related (ATG) proteins in yeast, and many of these proteins are highly conserved in mammals [76,77].
The initiation and expansion of the phagophore depend on the activation of the ULK1 and PI3-K complexes and the synthesis of phosphatidylinositol 3-phosphate (PI3P). In addition, ATG9-containing vesicles also provide additional lipid/membrane sources for pre-autophagosomal structures [78,79,80]. The elongation and the maturation of autophagosomes require the ATG12 conjugation system which includes ATG5, ATG7, ATG10, ATG12, and ATG16L1 proteins [81,82,83,84]. The covalent conjugation ATG12 and ATG5 is accomplished by two enzymes ATG7 and ATG10. The E1-like enzyme ATG7 activates ATG12 via the formation of a thioester bond between the C-terminus of ATG12 and the cysteine507 of ATG7 [85]. Subsequently, activated ATG12 is transferred to the E2 enzyme ATG10 [86], and is eventually conjugated to lys149 of ATG5 [81]. The conjugate ATG12-ATG5 subsequently forms a protein complex with ATG16L1 localized on the pre-autophagosomal structure [87]. The ATG12-ATG5-ATG16L complex exhibits E3-like activity that facilitates the conjugation of LC3 to phosphatidylethanolamine (PE). LC3 exists in two forms: the free mature form referred to LC3-I present in the cytosol and the rapidly lipidated form (LC3-II), which is formed during autophagosome formation [88]. LC3 is first processed by autophagin (ATG4) into LC3-I with the exposure of a glycine residue at the C-terminus [89]. The second conjugation system ensures that the cytosolic form of LC3 (LC3-I) links to lipid-conjugated phosphatidylethanolamine (PE) to form LC3-PE (LC3-II). The LC3-II protein is rapidly degraded when the autophagosome fuses with the lysosome. It has also been postulated that WD Repeat Domain, Phosphoinositide Interacting 2 (WIPI2) could act upstream of Atg16L1 to recruit the ATG12-ATG5-ATG16L1 and LC3 proteins to the autophagosome membrane [90]. Since LC3 localizes to the autophagosome membrane, monitoring the free mature form referred to LC3B-I present in the cytosol and the rapidly lipidated form (LCB3-II) by immunoblotting is currently the most widely used method to indirectly assess the levels of autophagosome formation [88]. Importantly, an increase in LC3B-II protein content may not necessarily imply that autophagy is elevated, since insufficient autophagosome fusion with lysosomes or defective lysosome function may also cause higher levels of the LC3B-II protein, irrespective of the degree of autophagy. Currently, autophagic flux assays are the gold standard method for assessing autophagy in a given cell or tissue [91]. In mouse skeletal muscle, autophagy flux has been accomplished by using inhibitors of autophagy, including for instance colchicine [36,92,93,94,95,96], leupeptin [92], and chloroquine [97]. In addition, GFP-LC3 transgenic mice are also often used to evaluate autophagy in vivo [98]. It is worth mentioning that monitoring autophagy in humans in vivo is impossible and that studies in humans are restricted to assessing indirect markers of autophagy activity. Extensive descriptions of autophagy monitoring procedures, as well as their advantages and limitations, have been reviewed in [91,99,100].
It is now generally accepted that autophagy contributes to mammalian cellular homoeostasis by removing damaged organelles under normal physiological conditions and even more under stress conditions, including nutrient and energy deprivation. For example, during acute starvation, autophagic production of amino acids in the liver is important for regulating hepatic gluconeogenesis [101]. There is also evidence that autophagy plays important roles in lifespan and health span in mammals [102,103,104,105]. In mice, muscle-specific deletion of ATG7 (a key regulator of autophagy) was shown to result in severe muscle atrophy, in the deterioration of neuromuscular junctions, and in a reduced lifespan [102]. This severe phenotype is associated with significant mitochondrial abnormalities and multilamellar body structures in mice with muscle-selective inhibition of autophagy [102,106,107,108]. Importantly, plasmid mediated-ATG7 overexpression for 14 days in aged skeletal muscles restored autophagy, ameliorated the integrity of neuromuscular junctions, and increased myofiber size [102]. In humans, mutations in ATG7 cause neurodevelopmental disorders involving neurologic, muscular, and endocrine hypofunction [109]. It was also reported that p62/SQSTM1 (sequestosome-1) accumulates in aged skeletal muscles, suggesting that a decline in autophagy may contribute to sarcopenia [110]. In line with this view, several studies have observed an accumulation of lipofuscin particles in aged muscles from rodents [47] and humans [111,112], suggesting that the autophagy-lysosomal function is defective. Furthermore, inactivation of autophagy regulating genes has also been associated with decreased lifespan in Caenorhabditis elegans [113], whereas promotion of autophagy in Drosophila extended lifespan [69,114]. It has also been reported that muscle-specific overexpression of FoxO transcription factors in Drosophila decreased muscle aging and extended life span, at least in part by sustaining basal levels of autophagy [115]. It is worth mentioning that FoxO transcription factors are now known to be major regulators of several autophagy-related proteins [116,117,118]. Transgenic overexpression of Atg5 (key regulator of autophagy) in mice was shown to improve motor function and extend lifespan by 17%, which is most likely due to enhanced autophagy [119]. There is evidence that aging-related alterations in autophagy likely lead to accumulation of damaged macromolecules and organelles during aging [102,120]. However, it is important to mention here that the molecular machinery regulating autophagy is still only partly understood. Further studies will, therefore, be required to refine our understanding of the exact role that autophagy plays in the skeletal muscle aging process.

5. The Role of Mitophagy in Skeletal Muscle Health and Aging

Mitophagy is a specific form of autophagy where damaged mitochondria are selectively eliminated by the autophagy-lysosome system. In healthy skeletal muscles, damaged and depolarized mitochondria are selectively removed through mitophagy pathways (Figure 3). The most extensively studied mitophagy pathway involves the E3 ligase Parkin and PINK1 (Phosphatase and tensin homolog (PTEN)-induced kinase 1) [121,122,123,124,125]. Parkin, an E3 ubiquitin ligase encoded by the Park2 gene, is a 465 amino acid protein that translocates to depolarized mitochondria to initiate mitophagy [121,122,123,124,125]. Loss of mitochondrial membrane potential (ΔΨm) by mitochondrial uncouplers blocks the import of PINK1 into the inner mitochondrial membrane (IMM). PINK1 then accumulates on the outer mitochondrial membrane (OMM) [122], which, in turn, stimulates the recruitment of Parkin E3 ligase activity from the cytosol followed by removal of damaged mitochondria by autophagosomes [124,125]. Once activated, Parkin has been reported to facilitate ubiquitination of numerous OMM proteins, including Mfn1 [126,127,128], Mfn2 [129], VDAC [130], Tom20 (translocase of the outer membrane 20) [121], and Miro [131]. These ubiquitinated proteins are then degraded by the proteasome and autophagy systems. In healthy mitochondria, cytosolic PINK1 protein is imported into the inner mitochondrial membrane and then degraded by mitochondrial proteases [132,133,134]. Importantly, PINK1 was recently shown to be dispensable for mitochondrial recruitment of Parkin and activation of mitophagy in cardiomyocytes [135]. It should be emphasized here that our current knowledge and understanding of Parkin-dependent mitophagy is largely based on cultured cell experiments in which non-physiological uncouplers such as CCCP were used to induce mitochondrial depolarization.
Mitophagy can also occur through Parkin-independent pathways. For instance, the selective elimination of damaged mitochondria can be achieved by p62/SQSTM1 [130], Bnip3 (BCL2 Interacting Protein 3), and Bnip3L (BCL2 Interacting Protein 3 Like, also called NIX) [136,137,138,139,140]. Moreover, it has been shown that mitochondrial ubiquitin chains can interact with optineurin (Optn) and NDP52 (Nuclear dot protein 52 kD) via its LC3 interaction region (LIR) motif to remove dysfunctional mitochondria [141,142]. A recent study also showed that deleting FUNDC1 (FUN14 Domain Containing 1) specifically in skeletal muscles in mice causes mitophagy impairment [143]. As such, when assessing mitophagy, it important to keep in mind that various proteins, such as Bnip3, Bnip3L, NDP52, and p62/SQSTM1, can also trigger mitophagy in a Parkin-independent manner (Figure 3). It should also be mentioned that the translocation of cardiolipin from the inner to outer mitochondrial membrane is another event that can trigger mitophagy [144].
Only few studies have examined the impact of aging on mitophagy pathways in skeletal muscle. Although conflicting data exist [54,145,146,147], several studies reported that mitophagy-related proteins are decreased in the skeletal muscle of old rodents [46,120,148,149,150,151,152]. Similarly, and while not a universal finding [54,145,146,147], Parkin content was reported lower in the skeletal muscle of old mice [152] and rats [151]. Importantly, the majority of studies that reported an increase in Parkin content in aged muscles also reported that mitophagy was impaired/insufficient in aged skeletal muscles [145,146,147]. In humans, we reported that the ratio of Parkin to VDAC (voltage-dependent anion channel) is significantly reduced in atrophied muscles of old men compared to muscles of young men, suggesting that mitophagy might be impaired during muscle aging in humans [9]. In line with this finding, it has been reported that inactive old women have significantly lower expression levels of Bnip3, Beclin 1, Atg7, and Park2 (the gene coding for Parkin), compared to a more active cohort [150]. These studies suggest the presence of a strong relationship between decreased markers of mitophagy and the development of skeletal muscle atrophy during aging. Based on the available literature, it is possible that disrupted mitophagy and unbalanced mitochondrial dynamics might lead to the accumulation of dysfunctional mitochondria during skeletal muscle aging (Figure 4).
Several recent studies using loss- and gain-of-function approaches in vivo have provided mechanistic insights into the physiological function of Parkin in skeletal muscle health and aging. For instance, it has been reported that muscles from Parkin-deficient mice display lower mitophagy flux in response to acute exercise [147] and endurance training [153]. We and others have reported that muscles of Parkin-deficient mice display several aging-like features, including contractile dysfunction [154] and impaired mitochondrial energetics [147,153,154,155]. We also reported that the loss of Parkin in skeletal muscle triggers an increase in the sensitization of the mPTP [154]. Taken together, these studies highlight the importance of Parkin in the maintenance of mitochondrial integrity in skeletal muscle. Using gain-of-function, we recently showed that Parkin overexpression increased muscle mass, fiber size, and mitochondrial enzyme activity in both young and old muscles [152]. Furthermore, in old mice, we found that Parkin overexpression increased muscle strength, increased mitochondrial content, and protected against oxidative stress, fibrosis, and apoptosis [152]. In line with these findings, Parkin overexpression in skeletal muscles of Drosophila led to higher mitochondrial citrate synthase enzymatic activity and attenuated the accumulation of protein aggregates, a marker of cellular aging [156], whereas Parkin loss-of-function in Drosophila resulted in a decreased lifespan and increased locomotor defects [157,158]. We also recently showed that Parkin overexpression prevents sepsis-induced skeletal muscle atrophy, likely by improving mitochondrial quality and content [159]. Taken altogether, these studies highlight the importance of proteins regulating mitophagy in skeletal muscle health and aging.
Importantly, mechanistic studies aiming at assessing the role played by proteins involved in Parkin-independent mitophagy in sarcopenia are currently lacking. Assessing whether modulating proteins regulating Parkin-independent mitophagy in aged mice can attenuate sarcopenia might open new therapeutic avenues.

6. Enhancing Mitophagy to Counteract Sarcopenia: Potential Applicability and Future Research

In the last years, several studies have provided the proof of principle that stimulating mitophagy can counteract sarcopenia. As discussed above, we have shown that overexpressing Parkin in aged skeletal muscles attenuates sarcopenia [152]. Recently, Dr Auwerx’s research group reported that urolithin A supplementation, a natural dietary microflora-derived metabolite from ingested ellagitannins and ellagic acid, stimulates Parkin-mediated mitophagy and prolongs lifespan in C. elegans [160], and improves muscle function and running performance in old rodents while increasing Parkin expression [160]. Administration of urolithin A in the first-in-human clinical trial has also been shown to be safe and to induce a molecular signature of mitochondrial health in the skeletal muscle of healthy sedentary elderly individuals [161]. Another natural compound from tomatoes, tomatidine, has also been shown to attenuate disuse-induced muscle atrophy and improve skeletal muscle mass and function in mice [162], and to prolong lifespan and health span in C. elegans by stimulating mitophagy [163]. It was also reported that lifespan extension in response to resveratrol supplementation in rats was associated with an activation of the Pink1/Parkin pathway [164]. It is worth mentioning that two small molecules, T0466 and T0467, were recently shown to increase the mitochondrial translocation of Parkin in human dopaminergic neurons and myoblasts [165]. Interestingly, both compounds were shown to reduce unfolded mitochondrial protein levels, presumably through enhanced PINK1-Parkin signaling [165]. Recently, it was shown that injecting old mice with a miR-181a mimic (miR-181a being a microRNA known to enhance mitophagy in myoblasts) resulted in improved skeletal muscle force, increased myofiber size, and a trend for increased succinate dehydrogenase activity [145]. Taken together, the available literature position mitophagy-modulating compounds as potential therapeutic tools to counteract skeletal muscle atrophy and weakness occurring with aging. Large-scale clinical trials are now needed to assess the efficacy of mitophagy-inducing compounds in fighting sarcopenia.
Interestingly, physical activity, which is well known to improve muscle and mitochondrial health in the elderly [166], was recently shown to stimulate mitophagy [57,153,167,168] and increase Parkin expression in the skeletal muscles of rodents [153] and humans [57,168]. It was also shown that physically active older adults display higher Parkin in mitochondrial fractions prepared from muscle biopsies. It is, therefore, possible that the beneficial effects of physical activity on muscle and mitochondrial health in older individuals [166] might, at least in part, involve an upregulation of mitophagy. Similarly, accumulating experimental evidence indicates that calorie restriction, one of the most efficient interventions to attenuate sarcopenia in rodents [25,26,27] and to improve mitochondrial health [28], also stimulates mitophagy (see [169] for a detailed review). Perhaps not surprisingly, evidence also indicates that fasting likely triggers mitophagy [169]. Interestingly, time-restricted eating, which involves regular periods of fasting, has emerged in the last few years as an intervention with the potential to improve mitochondrial health and increase health span [170,171]. It was even recently suggested that time-restricted eating might be an avenue to fight sarcopenia [172]. Whether time-restricted eating can efficiently stimulate mitophagy in aged skeletal muscles and attenuate the aging-related decline in muscle mass and function represent an interesting research avenue that requires further study.
Recently, it was reported that long-term exposure to a ketogenic diet attenuated the aging-related decline of relative muscle mass [173] and increased markers of mitochondrial content in aged skeletal muscles [174]. Interestingly, the ketone body β-hydroxybutyrate, which is elevated during nutritional ketosis, was shown to induce mitophagy in young and aged myocytes [175] and elicits favorable mitochondrial adaptations in myocytes [176]. These findings combined are particularly interesting when considering the availability of recently developed ketone esters known to significantly increase circulating β-hydroxybutyrate in humans [177]. However, whether a ketogenic diet or ketone supplements can efficiently stimulate mitophagy and attenuate sarcopenia in humans remains unknown.

7. Conclusions

In this review, we critically summarized the current state of knowledge on the role played by mitochondrial dynamics and mitophagy in skeletal muscle health and in the development of sarcopenia. Data from mechanistic studies have clearly highlighted that mitochondrial dynamics and mitophagy are critical to skeletal muscle and mitochondrial health. Whether impairment in mitochondrial dynamics contributes to sarcopenia still remains unknown, as the literature is filled with conflicting data. Studies aiming at modulating the expression of proteins regulating dynamics late in life should provide answers as to whether mitochondrial dynamics can be targeted to counter sarcopenia. While conflicting data also exist for mitophagy, a growing body of evidence indicates that mitophagy is impaired in aged skeletal muscles. Importantly, enhancing mitophagy through genetic or nutritional approaches improves skeletal muscle function in aged rodents and improves mitochondrial health in the elderly. Enhancing mitophagy in skeletal muscle, therefore, appears as a promising therapeutic target to prevent or even treat skeletal muscle dysfunction in the elderly.

Author Contributions

J.-P.L.-G., G.G., S.N.A.H. and E.B. designed and conceived this review. J.-P.L.-G. and G.G. prepared all figures and tables and wrote the first draft of the manuscript. S.N.A.H. and E.B. assisted in the writing and editing of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Natural Sciences and Engineering Council of Canada (NSERC) grant awarded to Gilles Gouspillou (RGPIN-2021-03724), by the Canadian Institutes of Health Research (CIHR) grants awarded to Sabah N.A. Hussain (MOP-93760) and Sabah N. A. Hussain and Gilles Gouspillou (MOV-409262), and by grants awarded to Esther Barreiro (FIS 18/00075 and FIS 21/00215 (FEDER), Spanish Ministry of Science and Innovation, Sociedad Española de Neumología y Cirugía Torácica (SEPAR) 2020 (Spain).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Gilles Gouspillou is supported by a Chercheur Boursier Junior 2 salary award from the FRQS. Jean-Philippe Leduc-Gaudet is supported by a FRQS postdoctoral fellowship. All figures were created with BioRender.com.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sartori, R.; Romanello, V.; Sandri, M. Mechanisms of muscle atrophy and hypertrophy: Implications in health and disease. Nat. Commun. 2021, 12, 330. [Google Scholar] [CrossRef]
  2. Xu, W.; Chen, T.; Cai, Y.; Hu, Y.; Fan, L.; Wu, C. Sarcopenia in Community-Dwelling Oldest Old Is Associated with Disability and Poor Physical Function. J. Nutr. Health Aging 2020, 24, 339–345. [Google Scholar] [CrossRef]
  3. Dos Santos, L.; Cyrino, E.S.; Antunes, M.; Santos, D.A.; Sardinha, L.B. Sarcopenia and physical independence in older adults: The independent and synergic role of muscle mass and muscle function. J. Cachexia Sarcopenia Muscle 2017, 8, 245–250. [Google Scholar] [CrossRef]
  4. Wahlen, B.M.; Mekkodathil, A.; Al-Thani, H.; El-Menyar, A. Impact of sarcopenia in trauma and surgical patient population: A literature review. Asian J. Surg. 2020, 43, 647–653. [Google Scholar] [CrossRef]
  5. Zhang, X.; Zhang, W.; Wang, C.; Tao, W.; Dou, Q.; Yang, Y. Sarcopenia as a predictor of hospitalization among older people: A systematic review and meta-analysis. BMC Geriatr. 2018, 18, 188. [Google Scholar] [CrossRef] [PubMed]
  6. Vetrano, D.L.; Landi, F.; Volpato, S.; Corsonello, A.; Meloni, E.; Bernabei, R.; Onder, G. Association of sarcopenia with short- and long-term mortality in older adults admitted to acute care wards: Results from the CRIME study. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 2014, 69, 1154–1161. [Google Scholar] [CrossRef] [Green Version]
  7. Anker, S.D.; Morley, J.E.; von Haehling, S. Welcome to the ICD-10 code for sarcopenia. J. Cachexia Sarcopenia Muscle 2016, 7, 512–514. [Google Scholar] [CrossRef] [PubMed]
  8. Larsson, L.; Degens, H.; Li, M.; Salviati, L.; Lee, Y.I.; Thompson, W.; Kirkland, J.L.; Sandri, M. Sarcopenia: Aging-Related Loss of Muscle Mass and Function. Physiol. Rev. 2019, 99, 427–511. [Google Scholar] [CrossRef]
  9. 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] [PubMed]
  10. Gouspillou, G.; Bourdel-Marchasson, I.; Rouland, R.; Calmettes, G.; Franconi, J.-M.; Deschodt-Arsac, V.; Diolez, P. Alteration of mitochondrial oxidative phosphorylation in aged skeletal muscle involves modification of adenine nucleotide translocator. Biochim. Biophys. Acta (BBA) Bioenerg. 2010, 1797, 143–151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Trounce, I.; Byrne, E.; Marzuki, S. Decline in skeletal muscle mitochondrial respiratory chain function: Possible factor in ageing. Lancet 1989, 1, 637–639. [Google Scholar] [CrossRef]
  12. Dirks, A.; Leeuwenburgh, C. Apoptosis in skeletal muscle with aging. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2002, 282, R519–R527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Chabi, B.; Ljubicic, V.; Menzies, K.J.; Huang, J.H.; Saleem, A.; Hood, D.A. Mitochondrial function and apoptotic susceptibility in aging skeletal muscle. Aging Cell 2008, 7, 2–12. [Google Scholar] [CrossRef]
  14. Short, K.R.; Bigelow, M.L.; Kahl, J.; Singh, R.; Coenen-Schimke, J.; Raghavakaimal, S.; Nair, K.S. Decline in skeletal muscle mitochondrial function with aging in humans. Proc. Natl. Acad. Sci. USA 2005, 102, 5618–5623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Wang, W.; Fang, H.; Groom, L.; Cheng, A.; Zhang, W.; Liu, J.; Wang, X.; Li, K.; Han, P.; Zheng, M.; et al. Superoxide flashes in single mitochondria. Cell 2008, 134, 279–290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Dodd, S.L.; Gagnon, B.J.; Senf, S.M.; Hain, B.A.; Judge, A.R. Ros-mediated activation of NF-kappaB and Foxo during muscle disuse. Muscle Nerve 2010, 41, 110–113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Sandri, M.; Sandri, C.; Gilbert, A.; Skurk, C.; Calabria, E.; Picard, A.; Walsh, K.; Schiaffino, S.; Lecker, S.H.; Goldberg, A.L. Foxo transcription factors induce the atrophy-related ubiquitin ligase atrogin-1 and cause skeletal muscle atrophy. Cell 2004, 117, 399–412. [Google Scholar] [CrossRef] [Green Version]
  18. Picca, A.; Lezza, A.M.S.; Leeuwenburgh, C.; Pesce, V.; Calvani, R.; Bossola, M.; Manes-Gravina, E.; Landi, F.; Bernabei, R.; Marzetti, E. Circulating Mitochondrial DNA at the Crossroads of Mitochondrial Dysfunction and Inflammation During Aging and Muscle Wasting Disorders. Rejuvenation Res. 2018, 21, 350–359. [Google Scholar] [CrossRef]
  19. Liu, Y.; Bi, X.; Zhang, Y.; Wang, Y.; Ding, W. Mitochondrial dysfunction/NLRP3 inflammasome axis contributes to angiotensin II-induced skeletal muscle wasting via PPAR-gamma. Lab. Investig. J. Tech. Methods Pathol. 2020, 100, 712–726. [Google Scholar] [CrossRef]
  20. Wang, X.H.; Zhang, L.; Mitch, W.E.; LeDoux, J.M.; Hu, J.; Du, J. Caspase-3 cleaves specific 19 S proteasome subunits in skeletal muscle stimulating proteasome activity. J. Biol. Chem. 2010, 285, 21249–21257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Coggan, A.R.; Spina, R.J.; King, D.S.; Rogers, M.A.; Brown, M.; Nemeth, P.M.; Holloszy, J.O. Skeletal muscle adaptations to endurance training in 60- to 70-yr-old men and women. J. Appl. Physiol. 1992, 72, 1780–1786. [Google Scholar] [CrossRef] [Green Version]
  22. Song, W.; Kwak, H.B.; Lawler, J.M. Exercise training attenuates age-induced changes in apoptotic signaling in rat skeletal muscle. Antioxid. Redox Signal. 2006, 8, 517–528. [Google Scholar] [CrossRef]
  23. St-Jean-Pelletier, F.; Pion, C.H.; Leduc-Gaudet, J.P.; Sgarioto, N.; Zovilé, I.; Barbat-Artigas, S.; Reynaud, O.; Alkaterji, F.; Lemieux, F.C.; Grenon, A. The impact of ageing, physical activity, and pre-frailty on skeletal muscle phenotype, mitochondrial content, and intramyocellular lipids in men. J. Cachexia Sarcopenia Muscle 2016. [Google Scholar] [CrossRef]
  24. Lozano-Montoya, I.; Correa-Pérez, A.; Abraha, I.; Soiza, R.L.; Cherubini, A.; O’Mahony, D.; Cruz-Jentoft, A.J. Nonpharmacological interventions to treat physical frailty and sarcopenia in older patients: A systematic overview—The SENATOR Project ONTOP Series. Clin. Interv. Aging 2017, 12, 721–740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Baker, D.J.; Betik, A.C.; Krause, D.J.; Hepple, R.T. No decline in skeletal muscle oxidative capacity with aging in long-term calorically restricted rats: Effects are independent of mitochondrial DNA integrity. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 2006, 61, 675–684. [Google Scholar] [CrossRef] [PubMed]
  26. Mayhew, M.; Renganathan, M.; Delbono, O. Effectiveness of caloric restriction in preventing age-related changes in rat skeletal muscle. Biochem. Biophys. Res. Commun. 1998, 251, 95–99. [Google Scholar] [CrossRef] [PubMed]
  27. Hepple, R.T.; Baker, D.J.; McConkey, M.; Murynka, T.; Norris, R. Caloric restriction protects mitochondrial function with aging in skeletal and cardiac muscles. Rejuvenation Res. 2006, 9, 219–222. [Google Scholar] [CrossRef] [PubMed]
  28. Gouspillou, G.; Hepple, R.T. Facts and controversies in our understanding of how caloric restriction impacts the mitochondrion. Exp. Gerontol. 2013, 48, 1075–1084. [Google Scholar] [CrossRef] [PubMed]
  29. Ogata, T.; Yamasaki, Y. Ultra-High-resolution scanning electron microscopy of mitochondria and sarcoplasmic reticulum arrangement in human red, white, and intermediate muscle fibers. Anat. Rec. 1997, 248, 214–223. [Google Scholar] [CrossRef]
  30. Chan, D.C. Fusion and fission: Interlinked processes critical for mitochondrial health. Annu. Rev. Genet. 2012, 46, 265–287. [Google Scholar] [CrossRef] [Green Version]
  31. Suen, D.F.; Norris, K.L.; Youle, R.J. Mitochondrial dynamics and apoptosis. Genes Dev. 2008, 22, 1577–1590. [Google Scholar] [CrossRef] [Green Version]
  32. Tilokani, L.; Nagashima, S.; Paupe, V.; Prudent, J. Mitochondrial dynamics: Overview of molecular mechanisms. Essays Biochem. 2018, 62, 341–360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Chen, H.; Vermulst, M.; Wang, Y.E.; Chomyn, A.; Prolla, T.A.; McCaffery, J.M.; Chan, D.C. Mitochondrial fusion is required for mtDNA stability in skeletal muscle and tolerance of mtDNA mutations. Cell 2010, 141, 280–289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Sebastian, D.; Sorianello, E.; Segales, J.; Irazoki, A.; Ruiz-Bonilla, V.; Sala, D.; Planet, E.; Berenguer-Llergo, A.; Munoz, J.P.; Sanchez-Feutrie, M.; et al. Mfn2 deficiency links age-related sarcopenia and impaired autophagy to activation of an adaptive mitophagy pathway. EMBO J. 2016, 35, 1677–1693. [Google Scholar] [CrossRef] [PubMed]
  35. Bell, M.B.; Bush, Z.; McGinnis, G.R.; Rowe, G.C. Adult skeletal muscle deletion of Mitofusin 1 and 2 impedes exercise performance and training capacity. J. Appl. Physiol. 2019, 126, 341–353. [Google Scholar] [CrossRef] [PubMed]
  36. Tezze, C.; Romanello, V.; Desbats, M.A.; Fadini, G.P.; Albiero, M.; Favaro, G.; Ciciliot, S.; Soriano, M.E.; Morbidoni, V.; Cerqua, C.; et al. Age-Associated Loss of OPA1 in Muscle Impacts Muscle Mass, Metabolic Homeostasis, Systemic Inflammation, and Epithelial Senescence. Cell Metab. 2017, 25, 1374–1389.e6. [Google Scholar] [CrossRef] [PubMed]
  37. Pereira, R.O.; Tadinada, S.M.; Zasadny, F.M.; Oliveira, K.J.; Pires, K.M.P.; Olvera, A.; Jeffers, J.; Souvenir, R.; Mcglauflin, R.; Seei, A.; et al. OPA1 deficiency promotes secretion of FGF21 from muscle that prevents obesity and insulin resistance. EMBO J. 2017, 36, 2126–2145. [Google Scholar] [CrossRef]
  38. Rodríguez-Nuevo, A.; Díaz-Ramos, A.; Noguera, E.; Díaz-Sáez, F.; Duran, X.; Muñoz, J.P.; Romero, M.; Plana, N.; Sebastián, D.; Tezze, C.; et al. Mitochondrial DNA and TLR9 drive muscle inflammation upon Opa1 deficiency. EMBO J. 2018, 37. [Google Scholar] [CrossRef]
  39. Touvier, T.; De Palma, C.; Rigamonti, E.; Scagliola, A.; Incerti, E.; Mazelin, L.; Thomas, J.L.; D’Antonio, M.; Politi, L.; Schaeffer, L.; et al. Muscle-specific Drp1 overexpression impairs skeletal muscle growth via translational attenuation. Cell Death Dis. 2015, 6, e1663. [Google Scholar] [CrossRef] [Green Version]
  40. Giovarelli, M.; Zecchini, S.; Martini, E.; Garrè, M.; Barozzi, S.; Ripolone, M.; Napoli, L.; Coazzoli, M.; Vantaggiato, C.; Roux-Biejat, P.; et al. Drp1 overexpression induces desmin disassembling and drives kinesin-1 activation promoting mitochondrial trafficking in skeletal muscle. Cell Death Differ. 2020, 27, 2383–2401. [Google Scholar] [CrossRef] [Green Version]
  41. Romanello, V.; Guadagnin, E.; Gomes, L.; Roder, I.; Sandri, C.; Petersen, Y.; Milan, G.; Masiero, E.; Del Piccolo, P.; Foretz, M.; et al. Mitochondrial fission and remodelling contributes to muscle atrophy. EMBO J. 2010, 29, 1774–1785. [Google Scholar] [CrossRef]
  42. Favaro, G.; Romanello, V.; Varanita, T.; Andrea Desbats, M.; Morbidoni, V.; Tezze, C.; Albiero, M.; Canato, M.; Gherardi, G.; De Stefani, D.; et al. DRP1-mediated mitochondrial shape controls calcium homeostasis and muscle mass. Nat. Commun. 2019, 10, 2576. [Google Scholar] [CrossRef] [PubMed]
  43. Dulac, M.; Leduc-Gaudet, J.P.; Reynaud, O.; Ayoub, M.B.; Guerin, A.; Finkelchtein, M.; Hussain, S.N.; Gouspillou, G. Drp1 knockdown induces severe muscle atrophy and remodelling, mitochondrial dysfunction, autophagy impairment and denervation. J. Physiol. 2020. [Google Scholar] [CrossRef] [PubMed]
  44. Avellaneda, J.; Rodier, C.; Daian, F.; Brouilly, N.; Rival, T.; Luis, N.M.; Schnorrer, F. Myofibril and mitochondria morphogenesis are coordinated by a mechanical feedback mechanism in muscle. Nat. Commun. 2021, 12, 2091. [Google Scholar] [CrossRef] [PubMed]
  45. Lee, T.T.; Chen, P.L.; Su, M.P.; Li, J.C.; Chang, Y.W.; Liu, R.W.; Juan, H.F.; Yang, J.M.; Chan, S.P.; Tsai, Y.C.; et al. Loss of Fis1 impairs proteostasis during skeletal muscle aging in Drosophila. Aging Cell 2021, 20, e13379. [Google Scholar] [CrossRef]
  46. Joseph, A.M.; Adhihetty, P.J.; Wawrzyniak, N.R.; Wohlgemuth, S.E.; Picca, A.; Kujoth, G.C.; Prolla, T.A.; Leeuwenburgh, C. Dysregulation of mitochondrial quality control processes contribute to sarcopenia in a mouse model of premature aging. PLoS ONE 2013, 8, e69327. [Google Scholar] [CrossRef] [Green Version]
  47. 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] [Green Version]
  48. O’Connell, K.; Ohlendieck, K. Proteomic DIGE analysis of the mitochondria-enriched fraction from aged rat skeletal muscle. Proteomics 2009, 9, 5509–5524. [Google Scholar] [CrossRef] [Green Version]
  49. Iqbal, S.; Ostojic, O.; Singh, K.; Joseph, A.M.; Hood, D.A. Expression of mitochondrial fission and fusion regulatory proteins in skeletal muscle during chronic use and disuse. Muscle Nerve 2013, 48, 963–970. [Google Scholar] [CrossRef]
  50. Faitg, J.; Leduc-Gaudet, J.P.; Reynaud, O.; Ferland, G.; Gaudreau, P.; Gouspillou, G. Effects of Aging and Caloric Restriction on Fiber Type Composition, Mitochondrial Morphology and Dynamics in Rat Oxidative and Glycolytic Muscles. Front. Physiol. 2019, 10, 420. [Google Scholar] [CrossRef] [Green Version]
  51. Koltai, E.; Hart, N.; Taylor, A.W.; Goto, S.; Ngo, J.K.; Davies, K.J.; Radak, Z. Age-Associated declines in mitochondrial biogenesis and protein quality control factors are minimized by exercise training. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2012, 303, R127–R134. [Google Scholar] [CrossRef] [Green Version]
  52. Leduc-Gaudet, J.P.; Picard, M.; St-Jean Pelletier, F.; Sgarioto, N.; Auger, M.J.; Vallee, 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] [Green Version]
  53. Capitanio, D.; Vasso, M.; De Palma, S.; Fania, C.; Torretta, E.; Cammarata, F.P.; Magnaghi, V.; Procacci, P.; Gelfi, C. Specific protein changes contribute to the differential muscle mass loss during ageing. Proteomics 2016, 16, 645–656. [Google Scholar] [CrossRef] [PubMed]
  54. Yeo, D.; Kang, C.; Gomez-Cabrera, M.C.; Vina, J.; Ji, L.L. Intensified mitophagy in skeletal muscle with aging is downregulated by PGC-1alpha overexpression in vivo. Free Radic. Biol. Med. 2019, 130, 361–368. [Google Scholar] [CrossRef]
  55. Zhao, L.; Zou, X.; Feng, Z.; Luo, C.; Liu, J.; Li, H.; Chang, L.; Wang, H.; Li, Y.; Long, J.; et al. Evidence for association of mitochondrial metabolism alteration with lipid accumulation in aging rats. Exp. Gerontol. 2014, 56, 3–12. [Google Scholar] [CrossRef] [PubMed]
  56. Konopka, A.R.; Suer, M.K.; Wolff, C.A.; Harber, M.P. Markers of Human Skeletal Muscle Mitochondrial Biogenesis and Quality Control: Effects of Age and Aerobic Exercise Training. J. Gerontol. Ser. A 2013, 69, 371–378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Balan, E.; Schwalm, C.; Naslain, D.; Nielens, H.; Francaux, M.; Deldicque, L. Regular Endurance Exercise Promotes Fission, Mitophagy, and Oxidative Phosphorylation in Human Skeletal Muscle Independently of Age. Front. Physiol. 2019, 10, 1088. [Google Scholar] [CrossRef] [Green Version]
  58. Distefano, G.; Standley, R.A.; Dubé, J.J.; Carnero, E.A.; Ritov, V.B.; Stefanovic-Racic, M.; Toledo, F.G.; Piva, S.R.; Goodpaster, B.H.; Coen, P.M. Chronological Age Does not Influence Ex-vivo Mitochondrial Respiration and Quality Control in Skeletal Muscle. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 2017, 72, 535–542. [Google Scholar] [CrossRef]
  59. Joseph, A.M.; Adhihetty, P.J.; Buford, T.W.; Wohlgemuth, S.E.; Lees, H.A.; Nguyen, L.M.; Aranda, J.M.; Sandesara, B.D.; Pahor, M.; Manini, T.M.; et al. The impact of aging on mitochondrial function and biogenesis pathways in skeletal muscle of sedentary high- and low-functioning elderly individuals. Aging Cell 2012, 11, 801–809. [Google Scholar] [CrossRef] [Green Version]
  60. Crane, J.D.; Devries, M.C.; Safdar, A.; Hamadeh, M.J.; Tarnopolsky, M.A. The effect of aging on human skeletal muscle mitochondrial and intramyocellular lipid ultrastructure. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 2010, 65, 119–128. [Google Scholar] [CrossRef] [Green Version]
  61. Wyckelsma, V.L.; Levinger, I.; McKenna, M.J.; Formosa, L.E.; Ryan, M.T.; Petersen, A.C.; Anderson, M.J.; Murphy, R.M. Preservation of skeletal muscle mitochondrial content in older adults: Relationship between mitochondria, fibre type and high-intensity exercise training. J. Physiol. 2017, 595, 3345–3359. [Google Scholar] [CrossRef] [Green Version]
  62. Pham, A.H.; McCaffery, J.M.; Chan, D.C. Mouse lines with photo-activatable mitochondria to study mitochondrial dynamics. Genesis 2012, 50, 833–843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Perry, C.G.; Lally, J.; Holloway, G.P.; Heigenhauser, G.J.; Bonen, A.; Spriet, L.L. Repeated transient mRNA bursts precede increases in transcriptional and mitochondrial proteins during training in human skeletal muscle. J. Physiol. 2010, 588, 4795–4810. [Google Scholar] [CrossRef]
  64. Granata, C.; Oliveira, R.S.F.; Little, J.P.; Renner, K.; Bishop, D.J. Training intensity modulates changes in PGC-1α and p53 protein content and mitochondrial respiration, but not markers of mitochondrial content in human skeletal muscle. FASEB J. 2016, 30, 959–970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Meinild Lundby, A.K.; Jacobs, R.A.; Gehrig, S.; de Leur, J.; Hauser, M.; Bonne, T.C.; Flück, D.; Dandanell, S.; Kirk, N.; Kaech, A.; et al. Exercise training increases skeletal muscle mitochondrial volume density by enlargement of existing mitochondria and not de novo biogenesis. Acta Physiol. 2018, 222. [Google Scholar] [CrossRef] [PubMed]
  66. Mishra, P.; Varuzhanyan, G.; Pham, A.H.; Chan, D.C. Mitochondrial Dynamics Is a Distinguishing Feature of Skeletal Muscle Fiber Types and Regulates Organellar Compartmentalization. Cell Metab. 2015, 22, 1033–1044. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Dahl, R.; Larsen, S.; Dohlmann, T.L.; Qvortrup, K.; Helge, J.W.; Dela, F.; Prats, C. Three-Dimensional reconstruction of the human skeletal muscle mitochondrial network as a tool to assess mitochondrial content and structural organization. Acta Physiol. 2015, 213, 145–155. [Google Scholar] [CrossRef]
  68. Lally, J.S.V.; Herbst, E.A.F.; Matravadia, S.; Maher, A.C.; Perry, C.G.R.; Ventura-Clapier, R.; Holloway, G.P. Over-Expressing Mitofusin-2 in Healthy Mature Mammalian Skeletal Muscle Does Not Alter Mitochondrial Bioenergetics. PLoS ONE 2013, 8, e55660. [Google Scholar] [CrossRef] [Green Version]
  69. 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] [PubMed]
  70. Twig, G.; Elorza, A.; Molina, A.J.; Mohamed, H.; Wikstrom, J.D.; Walzer, G.; Stiles, L.; Haigh, S.E.; Katz, S.; Las, G.; et al. Fission and selective fusion govern mitochondrial segregation and elimination by autophagy. EMBO J. 2008, 27, 433–446. [Google Scholar] [CrossRef] [Green Version]
  71. Dulac, M.; Leduc-Gaudet, J.P.; Cefis, M.; Ayoub, M.B.; Reynaud, O.; Shams, A.; Moamer, A.; Nery Ferreira, M.F.; Hussain, S.N.; Gouspillou, G. Regulation of muscle and mitochondrial health by the mitochondrial fission protein Drp1 in aged mice. J. Physiol. 2021. [Google Scholar] [CrossRef]
  72. Levine, B.; Mizushima, N.; Virgin, H.W. Autophagy in immunity and inflammation. Nature 2011, 469, 323–335. [Google Scholar] [CrossRef]
  73. Yorimitsu, T.; Klionsky, D.J. Autophagy: Molecular machinery for self-eating. Cell Death Differ. 2005, 12 (Suppl. 2), 1542–1552. [Google Scholar] [CrossRef] [Green Version]
  74. Rabinowitz, J.D.; White, E. Autophagy and metabolism. Science 2010, 330, 1344–1348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Levine, B.; Kroemer, G. Autophagy in the pathogenesis of disease. Cell 2008, 132, 27–42. [Google Scholar] [CrossRef] [Green Version]
  76. Yang, Z.; Klionsky, D.J. Eaten alive: A history of macroautophagy. Nat. Cell Biol. 2010, 12, 814–822. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Tsukada, M.; Ohsumi, Y. Isolation and characterization of autophagy-defective mutants of Saccharomyces cerevisiae. FEBS Lett. 1993, 333, 169–174. [Google Scholar] [CrossRef] [Green Version]
  78. Young, A.R.; Chan, E.Y.; Hu, X.W.; Kochl, R.; Crawshaw, S.G.; High, S.; Hailey, D.W.; Lippincott-Schwartz, J.; Tooze, S.A. Starvation and ULK1-dependent cycling of mammalian Atg9 between the TGN and endosomes. J. Cell Sci. 2006, 119, 3888–3900. [Google Scholar] [CrossRef] [Green Version]
  79. Yamada, T.; Carson, A.R.; Caniggia, I.; Umebayashi, K.; Yoshimori, T.; Nakabayashi, K.; Scherer, S.W. Endothelial nitric-oxide synthase antisense (NOS3AS) gene encodes an autophagy-related protein (APG9-like2) highly expressed in trophoblast. J. Biol. Chem. 2005, 280, 18283–18290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Orsi, A.; Razi, M.; Dooley, H.C.; Robinson, D.; Weston, A.E.; Collinson, L.M.; Tooze, S.A. Dynamic and transient interactions of Atg9 with autophagosomes, but not membrane integration, are required for autophagy. Mol. Biol. Cell 2012, 23, 1860–1873. [Google Scholar] [CrossRef] [PubMed]
  81. Mizushima, N.; Noda, T.; Yoshimori, T.; Tanaka, Y.; Ishii, T.; George, M.D.; Klionsky, D.J.; Ohsumi, M.; Ohsumi, Y. A protein conjugation system essential for autophagy. Nature 1998, 395, 395–398. [Google Scholar] [CrossRef]
  82. Mizushima, N.; Sugita, H.; Yoshimori, T.; Ohsumi, Y. A new protein conjugation system in human. The counterpart of the yeast Apg12p conjugation system essential for autophagy. J. Biol. Chem. 1998, 273, 33889–33892. [Google Scholar] [CrossRef] [Green Version]
  83. Mizushima, N.; Noda, T.; Ohsumi, Y. Apg16p is required for the function of the Apg12p-Apg5p conjugate in the yeast autophagy pathway. EMBO J. 1999, 18, 3888–3896. [Google Scholar] [CrossRef] [Green Version]
  84. Mizushima, N.; Kuma, A.; Kobayashi, Y.; Yamamoto, A.; Matsubae, M.; Takao, T.; Natsume, T.; Ohsumi, Y.; Yoshimori, T. Mouse Apg16L, a novel WD-repeat protein, targets to the autophagic isolation membrane with the Apg12-Apg5 conjugate. J. Cell Sci. 2003, 116, 1679–1688. [Google Scholar] [CrossRef] [Green Version]
  85. Tanida, I.; Mizushima, N.; Kiyooka, M.; Ohsumi, M.; Ueno, T.; Ohsumi, Y.; Kominami, E. Apg7p/Cvt2p: A novel protein-activating enzyme essential for autophagy. Mol. Biol. Cell 1999, 10, 1367–1379. [Google Scholar] [CrossRef] [Green Version]
  86. Shintani, T.; Mizushima, N.; Ogawa, Y.; Matsuura, A.; Noda, T.; Ohsumi, Y. Apg10p, a novel protein-conjugating enzyme essential for autophagy in yeast. EMBO J. 1999, 18, 5234–5241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Kuma, A.; Mizushima, N.; Ishihara, N.; Ohsumi, Y. Formation of the approximately 350-kDa Apg12-Apg5.Apg16 multimeric complex, mediated by Apg16 oligomerization, is essential for autophagy in yeast. J. Biol. Chem. 2002, 277, 18619–18625. [Google Scholar] [CrossRef] [Green Version]
  88. Kabeya, Y.; Mizushima, N.; Ueno, T.; Yamamoto, A.; Kirisako, T.; Noda, T.; Kominami, E.; Ohsumi, Y.; Yoshimori, T. LC3, a mammalian homologue of yeast Apg8p, is localized in autophagosome membranes after processing. EMBO J. 2000, 19, 5720–5728. [Google Scholar] [CrossRef]
  89. Kabeya, Y.; Mizushima, N.; Yamamoto, A.; Oshitani-Okamoto, S.; Ohsumi, Y.; Yoshimori, T. LC3, GABARAP and GATE16 localize to autophagosomal membrane depending on form-II formation. J. Cell Sci. 2004, 117, 2805–2812. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Dooley, H.C.; Razi, M.; Polson, H.E.; Girardin, S.E.; Wilson, M.I.; Tooze, S.A. WIPI2 links LC3 conjugation with PI3P, autophagosome formation, and pathogen clearance by recruiting Atg12-5-16L1. Mol. Cell 2014, 55, 238–252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Klionsky, D.J.; Abdel-Aziz, A.K.; Abdelfatah, S.; Abdellatif, M.; Abdoli, A.; Abel, S.; Abeliovich, H.; Abildgaard, M.H.; Abudu, Y.P.; Acevedo-Arozena, A.; et al. Guidelines for the use and interpretation of assays for monitoring autophagy (4th edition). Autophagy 2021, 1–382. [Google Scholar] [CrossRef]
  92. Mofarrahi, M.; Guo, Y.; Haspel, J.A.; Choi, A.M.; Davis, E.C.; Gouspillou, G.; Hepple, R.T.; Godin, R.; Burelle, Y.; Hussain, S.N. Autophagic flux and oxidative capacity of skeletal muscles during acute starvation. Autophagy 2013, 9, 1604–1620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Ju, J.-S.; Varadhachary, A.S.; Miller, S.E.; Weihl, C.C. Quantitation of “autophagic flux” in mature skeletal muscle. Autophagy 2010, 6, 929–935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Romanello, V.; Scalabrin, M.; Albiero, M.; Blaauw, B.; Scorrano, L.; Sandri, M. Inhibition of the Fission Machinery Mitigates OPA1 Impairment in Adult Skeletal Muscles. Cells 2019, 8, 597. [Google Scholar] [CrossRef] [Green Version]
  95. Huang, D.D.; Yan, X.L.; Fan, S.D.; Chen, X.Y.; Yan, J.Y.; Dong, Q.T.; Chen, W.Z.; Liu, N.X.; Chen, X.L.; Yu, Z. Nrf2 deficiency promotes the increasing trend of autophagy during aging in skeletal muscle: A potential mechanism for the development of sarcopenia. Aging 2020, 12, 5977–5991. [Google Scholar] [CrossRef] [PubMed]
  96. Oost, L.J.; Kustermann, M.; Armani, A.; Blaauw, B.; Romanello, V. Fibroblast growth factor 21 controls mitophagy and muscle mass. J. Cachexia Sarcopenia Muscle 2019, 10, 630–642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. De Palma, C.; Morisi, F.; Cheli, S.; Pambianco, S.; Cappello, V.; Vezzoli, M.; Rovere-Querini, P.; Moggio, M.; Ripolone, M.; Francolini, M.; et al. Autophagy as a new therapeutic target in Duchenne muscular dystrophy. Cell Death Dis. 2012, 3, e418. [Google Scholar] [CrossRef] [Green Version]
  98. Mizushima, N. Methods for monitoring autophagy using GFP-LC3 transgenic mice. Methods Enzymol. 2009, 452, 13–23. [Google Scholar] [CrossRef]
  99. Moulis, M.; Vindis, C. Methods for Measuring Autophagy in Mice. Cells 2017, 6, 14. [Google Scholar] [CrossRef] [Green Version]
  100. Martinet, W.; Roth, L.; De Meyer, G.R.Y. Standard Immunohistochemical Assays to Assess Autophagy in Mammalian Tissue. Cells 2017, 6, 17. [Google Scholar] [CrossRef] [Green Version]
  101. Schworer, C.M.; Mortimore, G.E. Glucagon-induced autophagy and proteolysis in rat liver: Mediation by selective deprivation of intracellular amino acids. Proc. Natl. Acad. Sci. USA 1979, 76, 3169–3173. [Google Scholar] [CrossRef] [Green Version]
  102. Carnio, S.; LoVerso, F.; Baraibar, M.A.; Longa, E.; Khan, M.M.; Maffei, M.; Reischl, M.; Canepari, M.; Loefler, S.; Kern, H.; et al. Autophagy impairment in muscle induces neuromuscular junction degeneration and precocious aging. Cell Rep. 2014, 8, 1509–1521. [Google Scholar] [CrossRef]
  103. Fernandez, A.F.; Sebti, S.; Wei, Y.; Zou, Z.; Shi, M.; McMillan, K.L.; He, C.; Ting, T.; Liu, Y.; Chiang, W.C.; et al. Disruption of the beclin 1-BCL2 autophagy regulatory complex promotes longevity in mice. Nature 2018, 558, 136–140. [Google Scholar] [CrossRef]
  104. Madeo, F.; Zimmermann, A.; Maiuri, M.C.; Kroemer, G. Essential role for autophagy in life span extension. J. Clin. Investig. 2015, 125, 85–93. [Google Scholar] [CrossRef] [Green Version]
  105. Melendez, A.; Talloczy, Z.; Seaman, M.; Eskelinen, E.L.; Hall, D.H.; Levine, B. Autophagy genes are essential for dauer development and life-span extension in C. elegans. Science 2003, 301, 1387–1391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Masiero, E.; Agatea, L.; Mammucari, C.; Blaauw, B.; Loro, E.; Komatsu, M.; Metzger, D.; Reggiani, C.; Schiaffino, S.; Sandri, M. Autophagy is required to maintain muscle mass. Cell Metab. 2009, 10, 507–515. [Google Scholar] [CrossRef]
  107. Lim, J.A.; Zare, H.; Puertollano, R.; Raben, N. Atg5(flox)-Derived Autophagy-Deficient Model of Pompe Disease: Does It Tell the Whole Story? Mol. Ther. Methods Clin. Dev. 2017, 7, 11–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Wu, J.J.; Quijano, C.; Chen, E.; Liu, H.; Cao, L.; Fergusson, M.M.; Rovira, I., II; Gutkind, S.; Daniels, M.P.; Komatsu, M.; et al. Mitochondrial dysfunction and oxidative stress mediate the physiological impairment induced by the disruption of autophagy. Aging 2009, 1, 425–437. [Google Scholar] [CrossRef] [Green Version]
  109. Collier, J.J.; Guissart, C.; Olahova, M.; Sasorith, S.; Piron-Prunier, F.; Suomi, F.; Zhang, D.; Martinez-Lopez, N.; Leboucq, N.; Bahr, A.; et al. Developmental Consequences of Defective ATG7-Mediated Autophagy in Humans. N. Engl. J. Med. 2021, 384, 2406–2417. [Google Scholar] [CrossRef]
  110. Sakuma, K.; Kinoshita, M.; Ito, Y.; Aizawa, M.; Aoi, W.; Yamaguchi, A. p62/SQSTM1 but not LC3 is accumulated in sarcopenic muscle of mice. J. Cachexia Sarcopenia Muscle 2016, 7, 204–212. [Google Scholar] [CrossRef] [PubMed]
  111. Hütter, E.; Skovbro, M.; Lener, B.; Prats, C.; Rabøl, R.; Dela, F.; Jansen-Dürr, P. Oxidative stress and mitochondrial impairment can be separated from lipofuscin accumulation in aged human skeletal muscle. Aging Cell 2007, 6, 245–256. [Google Scholar] [CrossRef]
  112. Renault, V.; Rolland, E.; Thornell, L.-E.; Mouly, V.; Butler-Browne, G. Distribution of satellite cells in the human vastus lateralis muscle during aging. Exp. Gerontol. 2002, 37, 1513–1514. [Google Scholar] [CrossRef]
  113. Toth, M.L.; Sigmond, T.; Borsos, E.; Barna, J.; Erdelyi, P.; Takacs-Vellai, K.; Orosz, L.; Kovacs, A.L.; Csikos, G.; Sass, M.; et al. Longevity pathways converge on autophagy genes to regulate life span in Caenorhabditis elegans. Autophagy 2008, 4, 330–338. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Aparicio, R.; Rana, A.; Walker, D.W. Upregulation of the Autophagy Adaptor p62/SQSTM1 Prolongs Health and Lifespan in Middle-Aged Drosophila. Cell Rep. 2019, 28, 1029–1040.e5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Demontis, F.; Perrimon, N. FOXO/4E-BP signaling in Drosophila muscles regulates organism-wide proteostasis during aging. Cell 2010, 143, 813–825. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Mammucari, C.; Milan, G.; Romanello, V.; Masiero, E.; Rudolf, R.; Del Piccolo, P.; Burden, S.J.; Di Lisi, R.; Sandri, C.; Zhao, J.; et al. FoxO3 controls autophagy in skeletal muscle in vivo. Cell Metab. 2007, 6, 458–471. [Google Scholar] [CrossRef]
  117. Zhao, J.; Brault, J.J.; Schild, A.; Cao, P.; Sandri, M.; Schiaffino, S.; Lecker, S.H.; Goldberg, A.L. FoxO3 coordinately activates protein degradation by the autophagic/lysosomal and proteasomal pathways in atrophying muscle cells. Cell Metab. 2007, 6, 472–483. [Google Scholar] [CrossRef] [Green Version]
  118. Milan, G.; Romanello, V.; Pescatore, F.; Armani, A.; Paik, J.-H.; Frasson, L.; Seydel, A.; Zhao, J.; Abraham, R.; Goldberg, A.L.; et al. Regulation of autophagy and the ubiquitin–proteasome system by the FoxO transcriptional network during muscle atrophy. Nat. Commun. 2015, 6, 6670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Pyo, J.O.; Yoo, S.M.; Ahn, H.H.; Nah, J.; Hong, S.H.; Kam, T.I.; Jung, S.; Jung, Y.K. Overexpression of Atg5 in mice activates autophagy and extends lifespan. Nat. Commun. 2013, 4, 2300. [Google Scholar] [CrossRef] [Green Version]
  120. Garcia-Prat, L.; Martinez-Vicente, M.; Perdiguero, E.; Ortet, L.; Rodriguez-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]
  121. 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]
  122. Narendra, D.P.; Jin, S.M.; Tanaka, A.; Suen, D.-F.; Gautier, C.A.; Shen, J.; Cookson, M.R.; Youle, R.J. PINK1 is selectively stabilized on impaired mitochondria to activate Parkin. PLoS Biol. 2010, 8, e1000298. [Google Scholar] [CrossRef] [Green Version]
  123. Kane, L.A.; Lazarou, M.; Fogel, A.I.; Li, Y.; Yamano, K.; Sarraf, S.A.; Banerjee, S.; Youle, R.J. PINK1 phosphorylates ubiquitin to activate Parkin E3 ubiquitin ligase activity. J. Cell Biol. 2014, 205, 143–153. [Google Scholar] [CrossRef]
  124. Vives-Bauza, C.; Zhou, C.; Huang, Y.; Cui, M.; de Vries, R.L.; Kim, J.; May, J.; Tocilescu, M.A.; Liu, W.; Ko, H.S.; et al. PINK1-dependent recruitment of Parkin to mitochondria in mitophagy. Proc. Natl. Acad. Sci. USA 2010, 107, 378–383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Matsuda, N.; Sato, S.; Shiba, K.; Okatsu, K.; Saisho, K.; Gautier, C.A.; Sou, Y.-S.; Saiki, S.; Kawajiri, S.; Sato, F. PINK1 stabilized by mitochondrial depolarization recruits Parkin to damaged mitochondria and activates latent Parkin for mitophagy. J. Cell Biol. 2010, 189, 211–221. [Google Scholar] [CrossRef] [PubMed]
  126. Glauser, L.; Sonnay, S.; Stafa, K.; Moore, D.J. Parkin promotes the ubiquitination and degradation of the mitochondrial fusion factor mitofusin 1. J. Neurochem. 2011, 118, 636–645. [Google Scholar] [CrossRef] [PubMed]
  127. Poole, A.C.; Thomas, R.E.; Yu, S.; Vincow, E.S.; Pallanck, L. The mitochondrial fusion-promoting factor mitofusin is a substrate of the PINK1/parkin pathway. PLoS ONE 2010, 5, e10054. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Ziviani, E.; Tao, R.N.; Whitworth, A.J. Drosophila parkin requires PINK1 for mitochondrial translocation and ubiquitinates mitofusin. Proc. Natl. Acad. Sci. USA 2010, 107, 5018–5023. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Chen, Y.; Dorn, G.W., II. PINK1-phosphorylated mitofusin 2 is a Parkin receptor for culling damaged mitochondria. Science 2013, 340, 471–475. [Google Scholar] [CrossRef] [Green Version]
  130. Geisler, S.; Holmström, K.M.; Skujat, D.; Fiesel, F.C.; Rothfuss, O.C.; Kahle, P.J.; Springer, W. PINK1/Parkin-mediated mitophagy is dependent on VDAC1 and p62/SQSTM1. Nat. Cell Biol. 2010, 12, 119–131. [Google Scholar] [CrossRef]
  131. Liu, S.; Sawada, T.; Lee, S.; Yu, W.; Silverio, G.; Alapatt, P.; Millan, I.; Shen, A.; Saxton, W.; Kanao, T.; et al. Parkinson’s disease-associated kinase PINK1 regulates Miro protein level and axonal transport of mitochondria. PLoS Genet. 2012, 8, e1002537. [Google Scholar] [CrossRef] [PubMed]
  132. Deas, E.; Plun-Favreau, H.; Gandhi, S.; Desmond, H.; Kjaer, S.; Loh, S.H.; Renton, A.E.; Harvey, R.J.; Whitworth, A.J.; Martins, L.M. PINK1 cleavage at position A103 by the mitochondrial protease PARL. Hum. Mol. Genet. 2011, 20, 867–879. [Google Scholar] [CrossRef] [PubMed]
  133. Greene, A.W.; Grenier, K.; Aguileta, M.A.; Muise, S.; Farazifard, R.; Haque, M.E.; McBride, H.M.; Park, D.S.; Fon, E.A. Mitochondrial processing peptidase regulates PINK1 processing, import and Parkin recruitment. EMBO Rep. 2012, 13, 378–385. [Google Scholar] [CrossRef]
  134. Jin, S.M.; Lazarou, M.; Wang, C.; Kane, L.A.; Narendra, D.P.; Youle, R.J. Mitochondrial membrane potential regulates PINK1 import and proteolytic destabilization by PARL. J. Cell Biol. 2010, 191, 933–942. [Google Scholar] [CrossRef] [Green Version]
  135. Kubli, D.A.; Cortez, M.Q.; Moyzis, A.G.; Najor, R.H.; Lee, Y.; Gustafsson, A.B. PINK1 Is Dispensable for Mitochondrial Recruitment of Parkin and Activation of Mitophagy in Cardiac Myocytes. PLoS ONE 2015, 10, e0130707. [Google Scholar] [CrossRef]
  136. Zhang, J.; Ney, P.A. Role of BNIP3 and NIX in cell death, autophagy, and mitophagy. Cell Death Differ. 2009, 16, 939–946. [Google Scholar] [CrossRef] [Green Version]
  137. Bellot, G.; Garcia-Medina, R.; Gounon, P.; Chiche, J.; Roux, D.; Pouysségur, J.; Mazure, N.M. Hypoxia-Induced Autophagy Is Mediated through Hypoxia-Inducible Factor Induction of BNIP3 and BNIP3L via Their BH3 Domains. Mol. Cell. Biol. 2009, 29, 2570–2581. [Google Scholar] [CrossRef] [Green Version]
  138. Hanna, R.A.; Quinsay, M.N.; Orogo, A.M.; Giang, K.; Rikka, S.; Gustafsson, A.B. Microtubule-associated protein 1 light chain 3 (LC3) interacts with Bnip3 protein to selectively remove endoplasmic reticulum and mitochondria via autophagy. J. Biol. Chem. 2012, 287, 19094–19104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Rikka, S.; Quinsay, M.N.; Thomas, R.L.; Kubli, D.A.; Zhang, X.; Murphy, A.N.; Gustafsson, A.B. Bnip3 impairs mitochondrial bioenergetics and stimulates mitochondrial turnover. Cell Death Differ. 2011, 18, 721–731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Novak, I.; Kirkin, V.; McEwan, D.G.; Zhang, J.; Wild, P.; Rozenknop, A.; Rogov, V.; Lohr, F.; Popovic, D.; Occhipinti, A.; et al. Nix is a selective autophagy receptor for mitochondrial clearance. EMBO Rep. 2010, 11, 45–51. [Google Scholar] [CrossRef] [Green Version]
  141. Wong, Y.C.; Holzbaur, E.L. Optineurin is an autophagy receptor for damaged mitochondria in parkin-mediated mitophagy that is disrupted by an ALS-linked mutation. Proc. Natl. Acad. Sci. USA 2014, 111, E4439–E4448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Heo, J.M.; Ordureau, A.; Paulo, J.A.; Rinehart, J.; Harper, J.W. The PINK1-PARKIN Mitochondrial Ubiquitylation Pathway Drives a Program of OPTN/NDP52 Recruitment and TBK1 Activation to Promote Mitophagy. Mol. Cell 2015, 60, 7–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Fu, T.; Xu, Z.; Liu, L.; Guo, Q.; Wu, H.; Liang, X.; Zhou, D.; Xiao, L.; Liu, L.; Liu, Y.; et al. Mitophagy Directs Muscle-Adipose Crosstalk to Alleviate Dietary Obesity. Cell Rep. 2018, 23, 1357–1372. [Google Scholar] [CrossRef] [PubMed]
  144. Chu, C.T.; Ji, J.; Dagda, R.K.; Jiang, J.F.; Tyurina, Y.Y.; Kapralov, A.A.; Tyurin, V.A.; Yanamala, N.; Shrivastava, I.H.; Mohammadyani, D.; et al. Cardiolipin externalization to the outer mitochondrial membrane acts as an elimination signal for mitophagy in neuronal cells. Nat. Cell Biol. 2013, 15, 1197–1205. [Google Scholar] [CrossRef] [Green Version]
  145. Goljanek-Whysall, K.; Soriano-Arroquia, A.; McCormick, R.; Chinda, C.; McDonagh, B. miR-181a regulates p62/SQSTM1, parkin, and protein DJ-1 promoting mitochondrial dynamics in skeletal muscle aging. Aging Cell 2020, 19, e13140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Carter, H.N.; Kim, Y.; Erlich, A.T.; Zarrin-Khat, D.; Hood, D.A. Autophagy and mitophagy flux in young and aged skeletal muscle following chronic contractile activity. J. Physiol. 2018, 596, 3567–3584. [Google Scholar] [CrossRef]
  147. Chen, C.C.W.; Erlich, A.T.; Crilly, M.J.; Hood, D.A. Parkin is required for exercise-induced mitophagy in muscle: Impact of aging. Am. J. Physiol. Endocrinol. Metab. 2018, 315, E404–E415. [Google Scholar] [CrossRef]
  148. McMullen, C.A.; Ferry, A.L.; Gamboa, J.L.; Andrade, F.H.; Dupont-Versteegden, E.E. Age-Related changes of cell death pathways in rat extraocular muscle. Exp. Gerontol. 2009, 44, 420–425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Gaugler, M.; Brown, A.; Merrell, E.; DiSanto-Rose, M.; Rathmacher, J.A.; Reynolds, T.H.T. PKB signaling and atrogene expression in skeletal muscle of aged mice. J. Appl. Physiol. 2011, 111, 192–199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Drummond, M.J.; Addison, O.; Brunker, L.; Hopkins, P.N.; McClain, D.A.; LaStayo, P.C.; Marcus, R.L. Downregulation of E3 ubiquitin ligases and mitophagy-related genes in skeletal muscle of physically inactive, frail older women: A cross-sectional comparison. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 2014, 69, 1040–1048. [Google Scholar] [CrossRef] [PubMed]
  151. Russ, D.W.; Wills, A.M.; Boyd, I.M.; Krause, J. Weakness, SR function and stress in gastrocnemius muscles of aged male rats. Exp. Gerontol. 2014, 50, 40–44. [Google Scholar] [CrossRef] [PubMed]
  152. Leduc-Gaudet, J.P.; Reynaud, O.; Hussain, S.N.; Gouspillou, G. Parkin overexpression protects from ageing-related loss of muscle mass and strength. J. Physiol. 2019, 597, 1975–1991. [Google Scholar] [CrossRef]
  153. Chen, C.C.W.; Erlich, A.T.; Hood, D.A. Role of Parkin and endurance training on mitochondrial turnover in skeletal muscle. Skelet Muscle 2018, 8, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Gouspillou, G.; Godin, R.; Piquereau, J.; Picard, M.; Mofarrahi, M.; Mathew, J.; Purves-Smith, F.M.; Sgarioto, N.; Hepple, R.T.; Burelle, Y.; et al. Protective role of Parkin in skeletal muscle contractile and mitochondrial function. J. Physiol. 2018, 596, 2565–2579. [Google Scholar] [CrossRef] [PubMed]
  155. Peker, N.; Donipadi, V.; Sharma, M.; McFarlane, C.; Kambadur, R. Loss of Parkin impairs mitochondrial function and leads to muscle atrophy. Am. J. Physiol. Cell Physiol. 2018, 315, C164–C185. [Google Scholar] [CrossRef] [Green Version]
  156. Rana, A.; Rera, M.; Walker, D.W. Parkin overexpression during aging reduces proteotoxicity, alters mitochondrial dynamics, and extends lifespan. Proc. Natl. Acad. Sci. USA 2013, 110, 8638–8643. [Google Scholar] [CrossRef] [Green Version]
  157. Greene, J.C.; Whitworth, A.J.; Kuo, I.; Andrews, L.A.; Feany, M.B.; Pallanck, L.J. Mitochondrial pathology and apoptotic muscle degeneration in Drosophila parkin mutants. Proc. Natl. Acad. Sci. USA 2003, 100, 4078–4083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Pesah, Y.; Pham, T.; Burgess, H.; Middlebrooks, B.; Verstreken, P.; Zhou, Y.; Harding, M.; Bellen, H.; Mardon, G. Drosophila parkin mutants have decreased mass and cell size and increased sensitivity to oxygen radical stress. Development 2004, 131, 2183–2194. [Google Scholar] [CrossRef] [Green Version]
  159. Leduc-Gaudet, J.P.; Mayaki, D.; Reynaud, O.; Broering, F.E.; Chaffer, T.J.; Hussain, S.N.A.; Gouspillou, G. Parkin Overexpression Attenuates Sepsis-Induced Muscle Wasting. Cells 2020, 9, 1454. [Google Scholar] [CrossRef]
  160. Ryu, D.; Mouchiroud, L.; Andreux, P.A.; Katsyuba, E.; Moullan, N.; Nicolet-Dit-Felix, A.A.; Williams, E.G.; Jha, P.; Lo Sasso, G.; Huzard, D.; et al. Urolithin A induces mitophagy and prolongs lifespan in C. elegans and increases muscle function in rodents. Nat. Med. 2016, 22, 879–888. [Google Scholar] [CrossRef]
  161. Andreux, P.A.; Blanco-Bose, W.; Ryu, D.; Burdet, F.; Ibberson, M.; Aebischer, P.; Auwerx, J.; Singh, A.; Rinsch, C. The mitophagy activator urolithin A is safe and induces a molecular signature of improved mitochondrial and cellular health in humans. Nat. Metab. 2019, 1, 595–603. [Google Scholar] [CrossRef] [PubMed]
  162. Dyle, M.C.; Ebert, S.M.; Cook, D.P.; Kunkel, S.D.; Fox, D.K.; Bongers, K.S.; Bullard, S.A.; Dierdorff, J.M.; Adams, C.M. Systems-based discovery of tomatidine as a natural small molecule inhibitor of skeletal muscle atrophy. J. Biol. Chem. 2014, 289, 14913–14924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Fang, E.F.; Waltz, T.B.; Kassahun, H.; Lu, Q.; Kerr, J.S.; Morevati, M.; Fivenson, E.M.; Wollman, B.N.; Marosi, K.; Wilson, M.A.; et al. Tomatidine enhances lifespan and healthspan in C. elegans through mitophagy induction via the SKN-1/Nrf2 pathway. Sci. Rep. 2017, 7, 46208. [Google Scholar] [CrossRef] [Green Version]
  164. Das, S.; Mitrovsky, G.; Vasanthi, H.R.; Das, D.K. Antiaging properties of a grape-derived antioxidant are regulated by mitochondrial balance of fusion and fission leading to mitophagy triggered by a signaling network of Sirt1-Sirt3-Foxo3-PINK1-PARKIN. Oxidative Med. Cell. Longev. 2014, 2014, 345105. [Google Scholar] [CrossRef]
  165. Shiba-Fukushima, K.; Inoshita, T.; Sano, O.; Iwata, H.; Ishikawa, K.-I.; Okano, H.; Akamatsu, W.; Imai, Y.; Hattori, N. A Cell-Based High-Throughput Screening Identified Two Compounds that Enhance PINK1-Parkin Signaling. iScience 2020, 23, 101048. [Google Scholar] [CrossRef]
  166. Cartee, G.D.; Hepple, R.T.; Bamman, M.M.; Zierath, J.R. Exercise Promotes Healthy Aging of Skeletal Muscle. Cell Metab. 2016, 23, 1034–1047. [Google Scholar] [CrossRef] [Green Version]
  167. Laker, R.C.; Drake, J.C.; Wilson, R.J.; Lira, V.A.; Lewellen, B.M.; Ryall, K.A.; Fisher, C.C.; Zhang, M.; Saucerman, J.J.; Goodyear, L.J.; et al. Ampk phosphorylation of Ulk1 is required for targeting of mitochondria to lysosomes in exercise-induced mitophagy. Nat. Commun. 2017, 8, 548. [Google Scholar] [CrossRef]
  168. Brandt, N.; Gunnarsson, T.P.; Bangsbo, J.; Pilegaard, H. Exercise and exercise training-induced increase in autophagy markers in human skeletal muscle. Physiol. Rep. 2018, 6, e13651. [Google Scholar] [CrossRef]
  169. Mehrabani, S.; Bagherniya, M.; Askari, G.; Read, M.I.; Sahebkar, A. The effect of fasting or calorie restriction on mitophagy induction: A literature review. J. Cachexia Sarcopenia Muscle 2020, 11, 1447–1458. [Google Scholar] [CrossRef] [PubMed]
  170. Longo, V.D.; Panda, S. Fasting, Circadian Rhythms, and Time-Restricted Feeding in Healthy Lifespan. Cell Metab. 2016, 23, 1048–1059. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Mattson, M.P.; Longo, V.D.; Harvie, M. Impact of intermittent fasting on health and disease processes. Ageing Res. Rev. 2017, 39, 46–58. [Google Scholar] [CrossRef] [PubMed]
  172. Tinsley, G.M.; Paoli, A. Time-Restricted eating and age-related muscle loss. Aging 2019, 11, 8741–8742. [Google Scholar] [CrossRef]
  173. Wallace, M.A.; Aguirre, N.W.; Marcotte, G.R.; Marshall, A.G.; Baehr, L.M.; Hughes, D.C.; Hamilton, K.L.; Roberts, M.N.; Lopez-Dominguez, J.A.; Miller, B.F.; et al. The ketogenic diet preserves skeletal muscle with aging in mice. Aging Cell 2021, 20, e13322. [Google Scholar] [CrossRef] [PubMed]
  174. Zhou, Z.; Hagopian, K.; Lopez-Dominguez, J.A.; Kim, K.; Jasoliya, M.; Roberts, M.N.; Cortopassi, G.A.; Showalter, M.R.; Roberts, B.S.; Gonzalez-Reyes, J.A.; et al. A ketogenic diet impacts markers of mitochondrial mass in a tissue specific manner in aged mice. Aging 2021, 13, 7914–7930. [Google Scholar] [CrossRef] [PubMed]
  175. Thai, P.N.; Seidlmayer, L.K.; Miller, C.; Ferrero, M.; Dorn, G.W., II; Schaefer, S.; Bers, D.M.; Dedkova, E.N. Mitochondrial Quality Control in Aging and Heart Failure: Influence of Ketone Bodies and Mitofusin-Stabilizing Peptides. Front. Physiol. 2019, 10, 382. [Google Scholar] [CrossRef] [Green Version]
  176. Parker, B.A.; Walton, C.M.; Carr, S.T.; Andrus, J.L.; Cheung, E.C.K.; Duplisea, M.J.; Wilson, E.K.; Draney, C.; Lathen, D.R.; Kenner, K.B.; et al. beta-Hydroxybutyrate Elicits Favorable Mitochondrial Changes in Skeletal Muscle. Int. J. Mol. Sci. 2018, 19, 2247. [Google Scholar] [CrossRef] [Green Version]
  177. Soto-Mota, A.; Vansant, H.; Evans, R.D.; Clarke, K. Safety and tolerability of sustained exogenous ketosis using ketone monoester drinks for 28 days in healthy adults. Regul. Toxicol. Pharmacol. RTP 2019, 109, 104506. [Google Scholar] [CrossRef]
Figure 1. Mitochondrial dynamics. Mitochondria are highly dynamic organelles that can change their morphology through fusion and fission events. Fusion results in mitochondrial elongation (top section), whereas fission fragments mitochondria (bottom section). Mitochondrial fusion is controlled by Mfn1 & Mfn2 on the outer mitochondrial membrane and by Opa1 on the inner mitochondrial membrane. For mitochondrial fission to occur, cytoplasmic Drp1 is recruited on the mitochondrial outer membrane, where it binds to its adaptors, Mff, MiD49, and MiD51, causing a ring- or spiral-like structure around mitochondrial constriction sites. Created with BioRender.com.
Figure 1. Mitochondrial dynamics. Mitochondria are highly dynamic organelles that can change their morphology through fusion and fission events. Fusion results in mitochondrial elongation (top section), whereas fission fragments mitochondria (bottom section). Mitochondrial fusion is controlled by Mfn1 & Mfn2 on the outer mitochondrial membrane and by Opa1 on the inner mitochondrial membrane. For mitochondrial fission to occur, cytoplasmic Drp1 is recruited on the mitochondrial outer membrane, where it binds to its adaptors, Mff, MiD49, and MiD51, causing a ring- or spiral-like structure around mitochondrial constriction sites. Created with BioRender.com.
Ijms 22 08179 g001
Figure 2. Schematic representation of the autophagy machinery. Autophagy consists of several stages: initiation, elongation and maturation, fusion to lysosome, and degradation and recycling. Autophagy is initiated by stress signals (top section). Both the Unc-51 Like Autophagy Activating Kinase (ULK1) and phosphatidylinositol-3 kinase (PI3-K) complexes are essential for initiation of the phagophore. ATG9 is involved in the delivery of lipids for the phagophore’s membrane expansion. The elongation and the maturation of the autophagosomal membrane are dependent on two ubiquitin-like conjugation systems, the ATG12 and microtubule-associated protein light chain 3 (LC3) conjugation systems. The first conjugation system requires ATG12, ATG5, ATG7, ATG10, and ATG16L1 proteins. This system works as E3-like enzymes for the second LC3 conjugation system. The LC3 conjugation system requires ATG4, ATG7, and ATG3 enzymes. Both conjugation systems are essential for the elongation and maturation of the autophagosome. The WD Repeat Domain, Phosphoinositide Interacting 2 (WIPI2) protein also promotes conjugation of LC3 with ATG16L1. Created with BioRender.com.
Figure 2. Schematic representation of the autophagy machinery. Autophagy consists of several stages: initiation, elongation and maturation, fusion to lysosome, and degradation and recycling. Autophagy is initiated by stress signals (top section). Both the Unc-51 Like Autophagy Activating Kinase (ULK1) and phosphatidylinositol-3 kinase (PI3-K) complexes are essential for initiation of the phagophore. ATG9 is involved in the delivery of lipids for the phagophore’s membrane expansion. The elongation and the maturation of the autophagosomal membrane are dependent on two ubiquitin-like conjugation systems, the ATG12 and microtubule-associated protein light chain 3 (LC3) conjugation systems. The first conjugation system requires ATG12, ATG5, ATG7, ATG10, and ATG16L1 proteins. This system works as E3-like enzymes for the second LC3 conjugation system. The LC3 conjugation system requires ATG4, ATG7, and ATG3 enzymes. Both conjugation systems are essential for the elongation and maturation of the autophagosome. The WD Repeat Domain, Phosphoinositide Interacting 2 (WIPI2) protein also promotes conjugation of LC3 with ATG16L1. Created with BioRender.com.
Ijms 22 08179 g002
Figure 3. Parkin-dependent and Parkin-independent mitophagy pathways. (A) Parkin-dependent mitophagy pathway: in dysfunctional mitochondria, PINK1 stabilizes on the OMM leading to the recruitment of Parkin through a series of modifications, including ubiquitination and phosphorylation of Parkin and ubiquitin. During mitophagy, Parkin polyubiquitinates various OMM proteins, including mitofusins (Mfn1 and Mfn2), VDAC, Tom20, and Miro. Ubiquitinated proteins are recognized by several adaptors, leading to the recruitment of an autophagosome and subsequent degradation of the mitochondrion. (B) Parkin-independent mitophagy pathway: mitochondrial proteins, such as Nix, Bnip3, FUNDC1, and NDP52, can act as mitophagy receptors, promoting the recruitment of the autophagosome around the cargo. Created with BioRender.com.
Figure 3. Parkin-dependent and Parkin-independent mitophagy pathways. (A) Parkin-dependent mitophagy pathway: in dysfunctional mitochondria, PINK1 stabilizes on the OMM leading to the recruitment of Parkin through a series of modifications, including ubiquitination and phosphorylation of Parkin and ubiquitin. During mitophagy, Parkin polyubiquitinates various OMM proteins, including mitofusins (Mfn1 and Mfn2), VDAC, Tom20, and Miro. Ubiquitinated proteins are recognized by several adaptors, leading to the recruitment of an autophagosome and subsequent degradation of the mitochondrion. (B) Parkin-independent mitophagy pathway: mitochondrial proteins, such as Nix, Bnip3, FUNDC1, and NDP52, can act as mitophagy receptors, promoting the recruitment of the autophagosome around the cargo. Created with BioRender.com.
Ijms 22 08179 g003
Figure 4. Conceptual model linking impairments in mitochondrial dynamics and mitophagy to the accumulation of mitochondria and the loss of muscle mass and function during aging. In healthy skeletal muscles, mitochondrial dynamics and mitophagy play central roles in maintaining mitochondrial integrity. The mitochondrial dynamics balance and mitophagy levels are represented by a black and red line, respectively; the blue line represents the level of mitochondrial dysfunction; the green line represents muscle mass and strength. Created with BioRender.com.
Figure 4. Conceptual model linking impairments in mitochondrial dynamics and mitophagy to the accumulation of mitochondria and the loss of muscle mass and function during aging. In healthy skeletal muscles, mitochondrial dynamics and mitophagy play central roles in maintaining mitochondrial integrity. The mitochondrial dynamics balance and mitophagy levels are represented by a black and red line, respectively; the blue line represents the level of mitochondrial dysfunction; the green line represents muscle mass and strength. Created with BioRender.com.
Ijms 22 08179 g004
Table 1. Compilation of studies that have investigated the impact of aging on the expression or content of proteins regulating mitochondrial dynamics.
Table 1. Compilation of studies that have investigated the impact of aging on the expression or content of proteins regulating mitochondrial dynamics.
SpeciesSexTissueModelMain ResultsRef.
MiceMale & female (combined)QUADYouths/Middle-aged
3–6 vs. 8–15 months old
Mfn1 & Mfn2: ↑
Opa1 & Drp1: NS
Fis1: ↓
[46]
Rats
(Fischer 344 Brown Norway)
MaleEDLYouths/Aged
5 vs. 35 months old
Mfn2, Fis1 & Opa1: ↑
Drp1: NS
[47]
Rats
(Wistar)
Not specifiedGASYouths/Aged
3 vs. 26 months old
Fis1: ↑[48]
Rats
(Fischer 344 Brown Norway)
MaleTAYouths/Aged
5 vs. 35 months old
Fis1 & Drp1;:↑
Mfn2: ↓
Opa1: NS
[49]
Rats
(Sprague–Dawley)
MaleGAS & SOLYouths/Aged
9 vs. 22 months old
Drp1 (SOL & GAS):↑
Mfn2 & Fis1 (GAS): ↑
[50]
Rats
(Wistar)
MaleGASYouths/Aged
3 vs. 26 months old
Fis1 & Mfn1: ↑ [51]
MiceMaleGASYouths/Aged
2–3 vs. 22–24 months old
Mfn2/Drp1 ratio:↑
Opa1, Drp1, Mfn1 & Mfn2: NS
[52]
Rats
(Sprague-Dawley)
MaleGAS & TRIYouths/Aged
3 vs. 22 months old
Opa1 & Mfn1 (GAS & TRI): ↑
Fis1 (GAS): ↓
Fis1 (TRI): ↑
[53]
MiceNot specifiedTAYouths/Aged
6 vs. 18 months old
OPA1: ↓[36]
MiceNot specifiedGASYouths/Aged
6 vs. 22 months old
Mfn1, Mfn2, Opa1 & Fis1: ↓
Drp1: NS
[34]
Mice
(C57BL/6J)
FemaleTAYouths/Aged
2 vs. 24 months old
Fis1 & Mfn2: ↑
Drp1 & Opa1: NS
[54]
Rats
(Sprague-Dawley)
MaleMusclesYouths/Aged
5 vs. 25 months old
Drp1: ↑
Opa1: NS
Mfn2 & Fis1: ↓
[55]
HumansMaleVLYounger men (20 ± 1 y) vs. older men (74 ± 3 y)Mfn1, Mfn2 & Fis1: NS[56]
HumansMaleVLYounger men (22 ± 1 y) vs. older men (67 ± 2 y)Opa1, Mfn2, Fis1: NS[57]
HumansMale & female (combined)VLYounger (24 ± 3 y) vs. older adults (78 ± 5 y)Opa1, Mfn2, Fis1 & Drp1: NS[58]
HumansMale & female (combined)VLYounger (23 ± 1 y) vs. older adults (75 ± 1 y)Mfn2, Fis1, Drp1: NS Opa1: ↓[59]
HumansMale & female (combined)VLYounger (≈23 y) vs. older adults (≈70 y)Mfn2: ↓
Drp1: trend for ↓
[60]
Abbreviations: Quadriceps (QUAD); Tibial anterior (TA); Soleus (SOL); Extensor digitorum longus (EDL); Vastus lateralis (VL) and Gastrocnemius (GAS); Triceps (TRI); ↓ indicate decreased levels, whereas ↑ indicate increased levels in aged skeletal muscles. No significant change (NS).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Leduc-Gaudet, J.-P.; Hussain, S.N.A.; Barreiro, E.; Gouspillou, G. Mitochondrial Dynamics and Mitophagy in Skeletal Muscle Health and Aging. Int. J. Mol. Sci. 2021, 22, 8179. https://doi.org/10.3390/ijms22158179

AMA Style

Leduc-Gaudet J-P, Hussain SNA, Barreiro E, Gouspillou G. Mitochondrial Dynamics and Mitophagy in Skeletal Muscle Health and Aging. International Journal of Molecular Sciences. 2021; 22(15):8179. https://doi.org/10.3390/ijms22158179

Chicago/Turabian Style

Leduc-Gaudet, Jean-Philippe, Sabah N. A. Hussain, Esther Barreiro, and Gilles Gouspillou. 2021. "Mitochondrial Dynamics and Mitophagy in Skeletal Muscle Health and Aging" International Journal of Molecular Sciences 22, no. 15: 8179. https://doi.org/10.3390/ijms22158179

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