Next Article in Journal
The Effects of Combined Exposure to Simulated Microgravity, Ionizing Radiation, and Cortisol on the In Vitro Wound Healing Process
Next Article in Special Issue
The Contribution of Tumor Derived Exosomes to Cancer Cachexia
Previous Article in Journal
A CD19-Anti-ErbB2 scFv Engager Protein Enables CD19-Specific CAR T Cells to Eradicate ErbB2+ Solid Cancer
Previous Article in Special Issue
The Role of Crosstalk between Adipose Cells and Myocytes in the Pathogenesis of Sarcopenic Obesity in the Elderly
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Age-Related Dysfunction in Proteostasis and Cellular Quality Control in the Development of Sarcopenia

by
Hector G. Paez
1,2,3,4,
Christopher R. Pitzer
1,2,3,4 and
Stephen E. Alway
1,2,3,4,5,*
1
Department of Physiology, College of Medicine, University of Tennessee Health Science Center, Memphis, TN 38163, USA
2
Integrated Biomedical Sciences Graduate Program, College of Graduate Health Sciences, University of Tennessee Health Science Center, Memphis, TN 38163, USA
3
Laboratory of Muscle Biology and Sarcopenia, Department of Physical Therapy, College of Health Professions, University of Tennessee Health Science Center, Memphis, TN 38163, USA
4
Center for Muscle, Metabolism and Neuropathology, Division of Regenerative and Rehabilitation Sciences, College of Health Professions, University of Tennessee Health Science Center, Memphis, TN 38163, USA
5
The Tennessee Institute of Regenerative Medicine, Memphis, TN 38163, USA
*
Author to whom correspondence should be addressed.
Cells 2023, 12(2), 249; https://doi.org/10.3390/cells12020249
Submission received: 10 December 2022 / Revised: 4 January 2023 / Accepted: 4 January 2023 / Published: 7 January 2023

Abstract

:
Sarcopenia is a debilitating skeletal muscle disease that accelerates in the last decades of life and is characterized by marked deficits in muscle strength, mass, quality, and metabolic health. The multifactorial causes of sarcopenia have proven difficult to treat and involve a complex interplay between environmental factors and intrinsic age-associated changes. It is generally accepted that sarcopenia results in a progressive loss of skeletal muscle function that exceeds the loss of mass, indicating that while loss of muscle mass is important, loss of muscle quality is the primary defect with advanced age. Furthermore, preclinical models have suggested that aged skeletal muscle exhibits defects in cellular quality control such as the degradation of damaged mitochondria. Recent evidence suggests that a dysregulation of proteostasis, an important regulator of cellular quality control, is a significant contributor to the aging-associated declines in muscle quality, function, and mass. Although skeletal muscle mammalian target of rapamycin complex 1 (mTORC1) plays a critical role in cellular control, including skeletal muscle hypertrophy, paradoxically, sustained activation of mTORC1 recapitulates several characteristics of sarcopenia. Pharmaceutical inhibition of mTORC1 as well as caloric restriction significantly improves muscle quality in aged animals, however, the mechanisms controlling cellular proteostasis are not fully known. This information is important for developing effective therapeutic strategies that mitigate or prevent sarcopenia and associated disability. This review identifies recent and historical understanding of the molecular mechanisms of proteostasis driving age-associated muscle loss and suggests potential therapeutic interventions to slow or prevent sarcopenia.

1. Introduction

Skeletal muscle comprises over a third of total body mass in young adults [1] and is a primary contributor to whole body metabolic rate and the overwhelming majority of postprandial glucose absorption [2]. Diminished muscle mass is associated with aging, and contributes to greater mortality and metabolic comorbidities [3,4]. Advancements in medical care in the last century have allowed for a remarkable extension of the human lifespan, despite the prevalence of chronic disease and disability in the last decades of life. In recent years, there has been a greater focus on extending the number of years lived in good health and not just increasing the life span per se. This has required an increased emphasis on the importance of muscle mass and quality in maintaining human health.
The concept of “healthy aging” and maintaining independence has been tied to optimizing muscle mass and strength. While muscle mass is critically important to the quality of life, aging is associated with a gradual reduction in muscle mass that plays a contributory role in the development of age-related diseases and disability [5,6]. Sarcopenia is a loss of skeletal muscle mass, quality, and function that is independent of weight loss and usually associated with aging. The European Working Group on Sarcopenia in Older People (EWGSOP) defines sarcopenia as a muscle disease resulting from a combination of low muscle mass and poor muscle function rooted in changes that accrue over a lifetime [7]. Loss of muscle mass has been shown to occur at a rate of 0.7–0.8% per year in the eighth decade of life; however, the loss of strength vastly outpaces the loss of muscle mass [8,9,10].
Sarcopenia is usually, but not always [11,12,13], associated with advanced age. The presence of low muscle mass and function associated with sarcopenia enhances the risk of falls and fractures [14,15], limits mobility [16,17], and increases the risk of disability [18], thus facilitating the necessity for long term assisted living [19]. Sarcopenia also represents a large fiscal burden on global healthcare institutions and government reimbursed healthcare, as patients with sarcopenia can incur larger healthcare cost [20,21]. While the disease sequalae of sarcopenia have been well characterized, the incidence of sarcopenia is expected to become more prominent with the ongoing demographic shift towards an older global population. Indeed, World Health Organization (WHO) projections predict that 1 in 6 people will be over the age of 60 by 2030 [22]. Low muscle mass and function are also central elements of frailty [23,24]. Frailty is an impaired resilience to disease that shares considerable overlap with the clinical manifestations of sarcopenia including an enhanced risk of mortality. Thus, understanding the mechanisms that regulate the loss of muscle mass and function are important if we are to reduce the effects of sarcopenia and frailty.
Well-established contributors to sarcopenia include neuromuscular remodeling [25,26,27], inflammation [28,29,30], apoptotic signaling [31,32,33], and alterations in the hormonal milieu [34,35,36], as well as environmental changes including physical inactivity [27,37,38] and nutritional inadequacies [39,40,41]. Proteostasis and cellular quality control mechanisms have garnered attention as possible points of cellular dysregulation in aging-associated disorders. Protein turnover is critical for tissue integrity, and requires a careful balance between anabolic and catabolic processes. Degradation of misfolded and damaged proteins is necessary to maintain proteome integrity and prevent organelle dysfunction and cellular degeneration. Furthermore, inadequate protein synthesis can result in a failure to replace degraded cellular components and compromised tissue maintenance. Initial lines of evidence supported a role for diminished protein synthesis in response to anabolic stimuli as a major contributor to sarcopenia. Although the mammalian target of rapamycin complex 1 (mTORC1) is critical for protein synthesis and skeletal muscle hypertrophy [42], recent evidence indicates that chronic mTORC1 activity can contribute to deficient cellular quality control and recapitulates several characteristics of sarcopenia including neuromuscular remodeling [43]. Our understanding of how dysregulated protein synthesis and protein degradation in aged skeletal muscle contributes to sarcopenia is incomplete. The present review summarizes previous and recent findings regarding the roles of protein synthesis, degradation, and cellular quality control mechanisms in sarcopenic muscle. Novel and recent studies that describe the interplay between anabolic and catabolic pathways are reviewed. Furthermore, we critically discuss the efficacy of caloric restriction as a treatment for sarcopenia and highlight challenges regarding the putative contributors to proteostasis dysfunction in the development of sarcopenia.

2. Skeletal Muscle Anabolism

2.1. Anabolic Signaling

Skeletal muscle retains remarkable plasticity by altering cellular size and contractile properties in response to altered stimuli. Muscle fiber size is established as the aggregate result of competition between concomitant rates of anabolic muscle protein synthesis (MPS) and catabolic muscle protein breakdown (MPB). Several anabolic stimuli can enhance MPS, including exercise, dietary amino acids, growth factors, and hormones. Among modulators of MPS, the serine/threonine kinase mammalian target of rapamycin (mTOR) is a critical component of mTORC1 that acts as a central node of regulation (Figure 1). Activation of mTORC1 is tightly regulated by the availability of nutrients and growth factors. The release of insulin by pancreatic β-cells during feeding initiates a cascade of signaling events in skeletal muscle that induces activation of the central effector of the phosphoinositide 3-kinase (PI3K) pathway, Ser/Thr kinase AKT (Akt), also known as protein kinase B (PKB). Insulin-mediated phosphorylation of Akt facilitates activation of mTORC1 through phosphorylation and inactivation of the tuberous sclerosis complex 1/2 (TSC1/2) [44], which acts as a GTPase-activating protein (GAP) for Ras homolog enriched in brain (Rheb) at the lysosomal surface and thereby prevents mTORC1 activation (Figure 1). Upon insulin-mediated phosphorylation by Akt, TSC1/2 dissociates from the lysosome and allows for Rheb activation of mTORC1 [45].
In addition to prompting the release of insulin, feeding also replenishes intracellular nutrients that can induce mTORC1 activation, in particular amino acids. Cytosolic availability of leucine is a potent signal for mTORC1 activation in large part through the leucine sensor Sestrin2. The availability of leucine reduces Sestrin2 activity, which relieves inhibition of the protein complex GTPase-activating protein activity toward Rags 2 (GATOR2) [46,47]. Uninhibited GATOR2 antagonizes the action of GTPase-activating protein activity toward Rags-1 (GATOR1), thereby relieving GATOR1 GAP activity on Rag guanosine triphosphatases [46] and allowing for their recruitment of mTORC1 to the lysosome in close proximity to Rheb. Activation of mTORC1 is also highly sensitive to arginine, which similarly to leucine, can relieve inhibition of GATOR2 through a cytosolic arginine sensor for mTORC1 subunit 1 (CASTOR1) dependent mechanism [48,49] (Figure 1). As such, a lysosome localized coincidence detector regulates cellular protein synthesis via mTORC1 by integrating extracellular growth signals with the intracellular nutrient status. Activation of mTORC1 inhibits degradative processes such as autophagy and stimulates mRNA translation in part through phosphorylation of targets eukaryotic translation initiation factor 4E (eIF4E)-binding protein 1 (4E-BP1), Eukaryotic elongation factor 2 (eEF2) kinase, and p70 ribosomal protein S6 kinase (p70S6K1). Phosphorylation of 4E-BP1 by mTORC1 results in the disassociation of 4E-BP1 from eIF4E and allows for the formation of the translation initiation complex at the 5′ end of mRNA to facilitate recruitment of the ribosomal machinery and cap dependent mRNA translation [50]. mTORC1 also increases protein synthesis through its enhancement of translation elongation via inhibitory phosphorylation of eEF2 kinase [51]. Activation of p70S6K by mTORC1 also boosts protein synthesis through augmentations in ribosomal biogenesis [52], translation initiation [53], and elongation [54].

2.2. Skeletal Muscle Anabolic Resistance

2.2.1. Amino Acid Induced MPS

Muscle atrophy occurs when myocytes have reduced protein accretion that results in reduced cellular size as the aggregate result of competition between reduced concomitant rates of MPS and increased MPB. Anabolic resistance refers to a blunted capacity to enhance MPS in response to an anabolic stimulus. Associations between aging and anabolic resistance have been under investigation for decades and have indicated that a perturbed skeletal muscle anabolic response to a nutritional stimulus exists in aged individuals [55,56]. While basal skeletal muscle fractional synthetic rates have been shown to be similar between young and aged individuals [56,57,58,59], the sensitivity and responsiveness of MPS have been reported to be reduced in the elderly in response to hyperaminoacidemia [56,57,60]. This suggests that senescent muscle may require a relatively high concentration of amino acids [61] to mount an adequate anabolic response. While the precise mechanism(s) that regulate the blunting of MPS in response to feeding in aging is/are not clear, a likely contributor is the reduction in physical activity that typically occurs with aging [62]. It is known that reduced physical activity can diminish the protein synthetic response to feeding [63] and this may play a large role the development of sarcopenia.
While there is significant support for an aging-suppression of MPS, not all studies have observed a blunted maximal MPS response to high doses of amino acids with elevated ages [64,65]. Furthermore, the concept that aging induces a reduced anabolic response to amino acid intake has also recently been challenged [66]. Nevertheless, while not a universal finding, the majority of studies support a role for attenuated MPS as a potential contributor to sarcopenia in aging.

2.2.2. MPS and Sarcopenic Obesity

The co-occurrence of age-related muscle loss and obesity, termed sarcopenic obesity, is a growing health concern. Sarcopenic obesity enhances the risk of developing type 2 diabetes mellitus and insulin resistance [37,67]. The elderly exhibit greater intramuscular lipid deposition than younger body mass and fat mass matched control subjects [68]. Aged animals also exhibit exacerbated insulin resistance when fed an obesogenic diet, which is concomitant with greater susceptibility to the accumulation of intramyocellular lipids [69,70]. Importantly, both lipid infusions and obesity have been shown to reduce the MPS response to amino acid and protein stimulation [71,72,73]. Investigations regarding the impact of coincident obesity and aging on the skeletal muscle protein synthetic response to protein ingestion have revealed that myofibrillar fractional synthetic rate is enhanced in young lean subjects and, to a significantly lesser degree, old lean subjects. However, old obese subjects fail to enhance post prandial myofibrillar fractional synthetic rate, indicating that obesity further dampens aging associated anabolic resistance to feeding [74]. Interestingly, while it is generally agreed upon that muscle loss with aging occurs at the expense of fast twitch type II fibers, it is not well understood if post prandial protein synthesis exhibits a fiber type specific deficit in aged muscle or how obesity may modulate this relationship. However, evidence exists to suggest that age-associated muscle atrophy may not be as restricted solely to fast twitch fibers, as was once thought to be the case in sarcopenia [75].

2.2.3. Gastrointestinal–Muscle Axis

Recent observations have linked the gut-muscle axis as a potential contributor and therapeutic target for aging associated anabolic resistance. While still in its infancy, several studies in rodents have linked alterations to the gut microbiota to aging and sarcopenia [76,77]. Indeed, the gut microbiome appears to have a major impact on skeletal muscle. For example, germ-free mice that are microbiologically sterile exhibit reduced skeletal muscle mass alongside diminished gene expression for oxidative metabolism. Importantly, these deficits are partially reversed by microbiota transplant from pathogen free mice [78]. Furthermore, gnotobiotic mice, which are germ-free mice transplanted with a known microbe, have demonstrated a causal link between specific microbes and skeletal muscle function. Gnotobiotic mice exhibited enhanced physical function both in the untrained state and after 4 weeks of exercise training when compared to germ-free mice, indicating an important role for gut microorganisms in physical function and exercise adaptations [79]. While mechanistic investigations regarding the interaction between gut microbes and skeletal muscle anabolism are sparse, analysis of the gastrointestinal microbiome of chickens revealed significant differences in microbe composition between chickens with low- and high-feed conversion ratios [80], a measure of the ratio of feed consumed to weight gained. These data imply that the gut microbiome composition can modulate the nutritive value of food consumed and the degree of anabolic growth.

2.2.4. Microbiome and Inflammation in Sarcopenia

It is interesting to note that a greater proportion of germ-free mice survived to old age than pathogen-free mice that are housed under standard laboratory conditions, which present with a more pronounced inflammatory state and enhanced gut permeability with increased age [81]. Germ free mice co-housed with aged pathogen-free mice exhibited a greater degree of intestinal permeability and circulating tumor necrosis factor α (TNFα) than mice that were co-housed with young pathogen free mice. This suggests that microbiome-associated changes with aging can influence the systemic inflammation of the host animal. Aging is also associated with reduced colonic mucosal barrier thickness [82] and a greater circulating concentrations of lipopolysaccharide [83]. Thus, it is possible that changes in the microbiome or gut health may be involved in heightened systemic inflammation associated with aging. The relationship between sarcopenia, aging, and chronic low-grade inflammation is well established and multiple lines of evidence have shown that chronic inflammation is deleterious to skeletal muscle health and increases protein turnover. Thus, alterations in the microbiome or gut function provide an attractive and plausible link to aging, inflammation, and skeletal muscle health.
It is important to note that while skeletal muscle is modulated by the microbiome, the gut microbiome can be modified by diet, antibiotics, and dietary supplements. The human microbiota has been reported to be similar between young and aged humans but there are larger differences between the young and the very old [84]. Furthermore, a study of long-term care and community dwelling seniors found that the diversity of the gut microbiome was significantly lower in individuals under long-term care when compared to similarly aged community dwelling seniors [85], highlighting the importance of considering environmental factors such as diet that may differ between individuals when investigating age-associated dysbiosis. These studies suggest that diet and age can affect skeletal muscle health both directly through the dietary supply of nutrients and indirectly through the dietary effect on the microbiome at any age. It is possible that alterations to the gut microbiome may also partly be due to environmental factors associated with frailty and sarcopenia in adults that may exhibit physical inactivity, alterations in diet, and consume medications for existing comorbidities.

2.3. Aging and Skeletal Muscle Recovery from Disuse

Skeletal muscle necessitates contractile and neural stimuli to maintain tissue integrity [86,87]. Extended periods of disuse such as bed rest, limb immobilization, and spaceflight can rapidly produce deficits in muscle mass. Muscle disuse induces a loss of contractile function leading to muscle atrophy, mitochondrial impairments, and functional decline [88,89]. The elderly are prone to sequential periods of disuse, due to more frequent falls and hospitalization than younger healthy subjects. In addition, the elderly exhibit poor regeneration of muscle tissue after an injury or disuse as compared to young subjects or animal models [90,91,92], which can potentially hasten the trajectory towards sarcopenia after repeated bouts of bed rest or immobilization. It is generally accepted that skeletal muscle atrophy during disuse occurs due to reductions in MPS and a heightened MPB [93], thereby compromising tissue maintenance. Importantly, aged animals show marked atrophy after disuse [94,95,96,97,98] and delayed restoration of contractile tissue [94,99,100,101], mitochondrial function, and dysfunctional β-oxidation after disuse compared to young animals [102]. Temporal assessments of the genetic response to hypertrophic stimuli have revealed metabolic gene expression to be perturbed by aging [103], which may play into the impaired response to reloading. Peroxisome proliferator-activated receptor-gamma coactivator (PGC)-1α/β-knockout mice have delayed force output recovery after disuse, indicating a crucial role for mitochondria in the functional recovery of skeletal muscle [104]. While protein ubiquitination and the expression of atrogenes muscle-specific RING finger protein 1 (MuRF1) and Atrogin1 are enhanced in aged animals during recovery after hindlimb suspension, proteosome activity and autophagic flux in some muscles are only enhanced in young animals in the reloading phase [105,106]. The accumulation of ubiquitinated proteins and reduced proteasomal activity in combination with an elevation of endoplasmic reticulum (ER) stress markers suggests dysregulated proteostasis in old animals [105], which could feasibly contribute to inadequate regeneration after disuse in geriatric muscle. Both protein synthesis and protein degradation are ATP-dependent processes. Thus, skeletal muscle remodeling is dependent upon adequate mitochondria number and function to supply ATP for protein synthesis and degradation. It has been estimated that the energetic cost of protein synthesis amounts to about 20% of basal ATP produced by oxidative phosphorylation [107,108]. While humans typically have enough fuel stores to sustain metabolic processes for several days to weeks, energy production is constrained by the rate of mitochondrial bioenergetics. While disuse impairs mitochondrial function [102,109], mitochondrial adaptations to exercise and activity increase mitochondrial number and function [110] and provide an optimal environment for muscle repair. Furthermore, muscle reloading in young animals reverses disuse-induced mitochondrial dysfunction [102,109]. In addition, both voluntary wheel running and a cocktail of mitochondrial targeted nutrients improve muscle regeneration after disuse [109,111], suggesting the restoration of mitochondrial function may facilitate tissue regrowth.
Regulators of mitochondrial volume and function are suppressed during muscle disuse as energy demands and contractile activity diminish. However, it is unclear if mitochondrial impairments during recovery may contribute to age-associated regenerative defects in skeletal muscle and contribute to sarcopenia. It is important to note that some studies have found that the amount of atrophy that occurs during disuse is not exacerbated by age [106,112] and the absolute extent of atrophy during disuse is sometimes more pronounced in young muscle [113], although this may not always be the case during short term periods of disuse [114]. This may be due in part because muscles in aged adult animals or humans are sarcopenic and have a smaller muscle mass and therefore have less protein to lose during disuse even if the rate of muscle loss is the same in young and old animals.
Another possibility is that the rate of MPS after disuse may be altered with aging. This is supported by observations that the larger muscles in young animals exhibit a more pronounced drop in MPS during disuse when compared to smaller muscles in aged animals [106]. When MPS was measured during recovery from disuse in aged animals, it was found that the rate of MPS was not diminished in 28-month-old when compared to 24-month-old rats, and in some muscles was significantly elevated with older age [106]. Furthermore, MPS was found to be elevated in old but not young rats during reloading despite blunted recovery [105]. If MPS is not impaired with aging, yet protein accretion is impaired in older hosts, then it is possible that geriatric muscle has a reduced ability to coordinate processes of protein synthesis, protein folding, and degradation of contractile tissue during reloading. A dysregulation of protein assembly in aging would ultimately compromise muscle remodeling and the reacquisition of muscle mass and this could contribute to sarcopenia if muscle mass and function are not restored after disuse.

2.4. Sarcopenia and mTORC1 Signaling

Although mTORC1 is an important regulator of muscle growth, recent evidence has implicated that sustained mTORC1 activation may be a potential contributor to age-associated myopathy (Figure 2). Despite reported insensitivity of MPS to external stimuli in aging, there are several reports suggesting that geriatric skeletal muscle exhibits signs of greater basal mTORC1 activity [43,115,116,117,118,119].
How alterations in senescent muscle drive greater basal mTORC1 activity or how sustained mTORC1 activity may differ between young and aged muscle have yet to be determined. Nevertheless, several studies have provided some clues to how mTORC1 could be elevated in aged muscle. For example, data from a transgenic mouse model that allows for the muscle-specific inducible expression of activated Akt has shown that activation of Akt (and presumably also mTORC1) for three weeks, results in greater muscle mass in young [120] and old (24 months) mice [115]. However, in contrast to young mice, old mice exhibited a reduction in force production when normalized to muscle mass (lower specific force), which was concomitant with morphological signs of fiber degeneration and immune cell infiltration [115]. These data imply that the chronic activation of the same anabolic signaling pathways (i.e., mTORC1) can produce very different effects in aged and young muscle. In a separate study, acute activation of mTORC1 in skeletal muscle by shRNA knockdown of TSC produced fiber hypertrophy [121], while sustained mTORC1 activation by genetic skeletal muscle TSC1-knockout, diminished muscle size and reduced total lean mass [121,122,123]. Indeed, inactivation of mTORC1 by muscle specific regulatory associated protein of mTOR (RAPTOR)-knockout also produces skeletal muscle atrophy [124]. Taken together, it appears that the proper balance of mTORC1 is required for optimal muscle health because both chronic loss and hyperactivation of mTORC1 can be deleterious for maintenance of skeletal muscle mass. The proposed model of mTORC1 dysregulation in aging leading to muscle atrophy is shown in Figure 2.
There is a complex regulation of muscle mass in aging because inhibition of chronically active mTORC1 with either rapamycin or low dose rapalogs attenuates aging-associated loss of muscle function and muscle mass in some, but not all, muscles [43,117,125]. Mechanistically, it is possible that hyperactivity of mTORC1 may negatively impact neuromuscular junction function, as both skeletal muscle specific TSC1-knockout mice and aged mice exhibit morphological alterations that are characteristic of neuromuscular junction instability. Importantly, neuromuscular deficits in mTORC1 hyperactive muscle and in aged mice are both partially reversed by mTORC1 inhibition via rapamycin administration [43]. Dysfunction of the motor neuron is a known contributor to the development of sarcopenia [126]. These findings support the idea that sustained mTORC1 activation in geriatric muscle can contribute to neuromuscular junction instability (Figure 2) and subsequently muscle atrophy and functional deficits, which may be partially reversed by rapamycin treatment. While rapamycin has been shown to produce favorable effects on sarcopenic muscle, it remains unclear if similarly inhibiting mTORC1 during reloading after disuse would enhance recovery and ameliorate ER stress in aged muscle (Figure 2).

3. Skeletal Muscle Catabolism

3.1. The Ubiquitin Proteasome System

3.1.1. Ubiquitin Proteasome Degradation of Cellular Proteins

Protein turnover is critical for cellular quality control under both basal and physiologically stressful conditions to maintain the integrity of the proteome. Degradation of cellular proteins occurs primarily through the ubiquitin proteasome system (UPS), autophagy, and to a lesser extent the Ca2+ dependent calpain system and caspase-3 [127,128]. Inadequate protein and organellar degradation can result in the accumulation of misfolded protein aggregates and dysfunctional cellular organelles. The primary component of the UPS is the 26S proteasome which is comprised of a core proteolytic 20S complex and the regulatory 19S complex capping one end of the 20S, although both ends of the core 20S can be capped to form the 30S proteasome [129]. The 19S regulatory particle is involved in the process of substrate engagement and subsequent ATP-dependent processive translocation and unfolding of the target polypeptide into the degradative 20S barrel subunit [130]. Proteins marked for degradation are recognized by the 26S proteasome through the presence of conjugated polyubiquitin tails.
Ubiquitination of damaged or misfolded proteins is a complex and highly regulated process involving a molecular hierarchy of enzymes involved in ubiquitin activation, conjugation, and ligation (reviewed in [130,131]). Briefly, conjugation of ubiquitin is first initiated by the ubiquitin-activating enzyme (E1), an ATP-dependent process that that results in a high energy thioester bond between E1 and ubiquitin. Secondly, there is the transfer of the ubiquitin-thioester to a ubiquitin conjugating protein (E2), which is followed by the conjugation of ubiquitin by E3 ubiquitin ligases to target proteins. While protein ubiquitination also serves other important non-proteasomal degradation roles such as, alterations of protein localization and function, these roles will not be covered in this review.
Relative to the small number of proteins within the human E1 (n = 2) and E2 (n = ~40) family, there are hundreds of different E3 ubiquitin ligases that confer specificity to the UPS, allowing for tight control of substrate ubiquitination [131]. Activation of the UPS has been implicated as a contributor to skeletal muscle atrophy under several pathological conditions including aging-induced sarcopenia [132,133], type 1 diabetes [134], muscle disuse [135,136,137,138,139], and fasting [138,140]. The two best-characterized E3 ubiquitin ligases implicated in skeletal muscle atrophy are Atrogin1 and MuRF1 [137]. Atrogin1 has been shown to mediate degradation of myogenic factors Myogenic Differentiation 1 (MyoD1) [141,142], myogenin [143], and eukaryotic translation initiation factor 3 subunit F (eIF3-f) [144]. Additionally, both Atrogin1 and MuRF1 have been shown to interact with sarcomeric proteins [145,146,147]. Furthermore, MuRF1 has also been implicated in the regulation of acetylcholine receptors at the neuromuscular junction [148]. Expression of atrogenes Atrogin1 and MuRF1 is regulated by Forkhead box O (FoxO) transcription factor translocation into the nucleus [138,149,150], although FoxO has also been shown to regulate autophagy [151]. Importantly, Akt can prevent FoxO nuclear localization [150], while FoxO activity is enhanced by energy deprivation through AMP-activated protein kinase (AMPK) [152,153].

3.1.2. The Ubiquitin Proteasome System in Sarcopenia

Given the ubiquitous FoxO-mediated expression of Atrogin1 and MuRF1 in a range of atrophic conditions, the expression of Atrogin1 and MuRF1 has often been used as a proxy measure for protein breakdown in skeletal muscle. However, it is important to note that enhanced expression of Atrogin1 and MuRF1, and even enhanced protein ubiquitination, is not synonymous with enhanced proteasome activity. Indeed, muscles from aged rats subjected to hindlimb unloading followed by reloading exhibit greater Atrogin1 and MuRF1 gene expression alongside elevated protein ubiquitination as compared to young adult animals; however, proteasome activity was generally unchanged or reduced throughout unloading and reloading in the older rodents [105]. These responses differed from muscles in young adult animals and the responses can even differ between muscles of older animals. This may represent a mismatch between ubiquitination and proteosome activity in geriatric muscle and potentially impaired UPS function during skeletal muscle reloading with aging [132,133].
The UPS system is tightly regulated to ensure proper protein turnover which is necessary to maintain a high quality of the muscle protein. Loss of protein quality control mechanisms is a hallmark of aging and a decline in protein turnover and the function of the UPS has long been thought to contribute to age-related cellular dysfunction [154]. The ability to rapidly degrade proteins in molecular signaling pathways is critical to avoid aberrant signaling and maintain proteome stability. Indeed, while overactivation of the UPS results in skeletal muscle atrophy, inadequate proteasome function in the skeletal muscle of Rpt3-knockout mice severely retards growth, resulting in reduced muscle mass, function, and signs of fiber degeneration concomitant with enhanced expression of Atrogin1, MuRF1, and protein ubiquitination [155]. The elevation in atrogene expression and protein ubiquitination may indicate a compensatory effort to ramp up protein degradation in the absence of adequate proteostasis. Similarly, MuRF1 and MuRF3 double-knockout mice exhibit myofiber protein aggregates, ultrastructural abnormalities, Z-line streaming, and pronounced functional deficits despite greater muscle mass [156]. Somewhat counterintuitively, MuRF1 null mice that are significantly protected from denervation induced atrophy also display a greater enhancement of skeletal muscle proteasome activity than wild type mice [157]. Similarly, while wildtype mice present a reduction in proteasome activity by 24 months of age, MuRF1 null mice exhibit enhanced proteasome activity and preserved skeletal muscle mass and fiber cross sectional area. Aged wildtype mice also showed signs of enhanced ER stress with age, which was attenuated in MuRF1 null mice [158]. Interestingly, despite a preservation of muscle mass, old MuRF1 null mice presented functional deficits in nerve-stimulated muscle force production when compared to age-matched wildtype mice. Given the role of MuRF1 in acetylcholine receptor regulation [148], it is possible that the observed defect may originate from inadequate neuromuscular remodeling.
Whether skeletal muscle UPS activity is enhanced or diminished with aging is controversial, with several studies pointing towards an increase [159,160,161], whereas others report a decrease of UPS activity in geriatric muscles [158,162,163]. Despite this lack of consensus, recent evidence has provided a surprising connection between elevated skeletal muscle mTORC1 signaling and activation of degradation pathways including the UPS system. Kaiser et al. [164] observed that mTORC1 hyperactivation in skeletal muscle of TSC1 null mice (a mouse model that recapitulates several characteristics of an aged phenotype), is concomitant with marked increases in several components of the UPS. Skeletal muscle from TSC1-knockout mice exhibit greater expression of several atrogenes and components of the 26S proteasome that is reversed by acute 3-day treatment with rapamycin. Remarkably, transcriptomic changes observed in muscles from TSC1-knockout mice after rapamycin treatment were highly similar to those observed in 30-month-old rapamycin treated mice [43,164]. Hyperactivation of mTORC1 in skeletal muscle of TSC1 muscle-knockout mice was concomitant with enhanced protein abundance of Nuclear Factor Erythroid 2-Like 1 (NRF1); a potent positive regulator of proteasomal gene expression. Knocking down NRF1 expression via shRNA reduced proteasomal protein abundance and catalytic activity, however this was insufficient to rescue muscle mass in TSC1 skeletal muscle-knockout mice [164].
Importantly, greater mTORC1 activity can impede AKT activation through p70S6K1-mediated phosphorylation of Insulin receptor substrate 1 (IRS-1) [165], thereby inhibiting downstream AKT activation. To test whether inhibition of AKT by hyperactive mTORC1 in TSC1-knockout mice was responsible for the apparent muscle atrophy, TSC1 muscle-knockout mice were bred with mutant AKT-TG mice that express a tamoxifen inducible form of active AKT. In line with AKT’s role as a negative regulator of FoxO and atrogene expression, AKT activation induced muscle growth in TSC1-knockout muscle. These data indicate that AKT-mediated suppression of FoxO targets, but not suppression of NRF1 and proteasomal gene expression, rescues muscle atrophy in the background of mTORC1 hyperactivation. As aging mice have been reported to have chronically high levels of mTORC1, as discussed above, the reversal of the suppression of AKT-regulated UPS by mTORC1 may provide a potential target for therapeutic reduction of sarcopenia in aging. However, simply increasing AKT may not be an adequate approach for treating sarcopenia because despite enhanced muscle size, sustained hyperactivation of AKT was reported to produce vacuolated fibers that are characteristic of late onset myopathy during mTORC1 hyperactivation. Nevertheless, the elegant work of Kaiser et al. [164] showed that proteasomal upregulation is an attempt to maintain proteostasis during mTORC1 hyperactivation (and diminished autophagy) and inhibition of FoxO-mediated gene expression results in early onset myopathy.

3.2. Autophagy and Mitophagy

3.2.1. Autophagy and Mitophagy Signaling

Similar to the UPS, autophagic degradation represents another major arm of protein quality control, to ensure cellular clearance of defective proteins and organelles as well as to recycle substrates during nutrient scarcity. A basal level of autophagy is important for cellular homeostasis, and the autophagic machinery is responsive to several stimuli including fasting [166,167] and exercise [168]. However, autophagy is elevated in response to oxidative stress [169], ER stress [170], and to pathological states such as cancer cachexia [171]. There exist three well-characterized types of autophagy which include microautophagy, chaperone-mediated autophagy, and macroautophagy. Macroautophagy primarily manages the clearance of cellular organelles and protein aggregates via engulfment by the autophagosome and degradation in lytic compartments through fusion with the lysosome. For the purposes of this review, macroautophagy will be referred to as autophagy (Figure 3).
The initiation of autophagy and autophagosome biogenesis is largely regulated by unc-51 like autophagy activating kinase 1 (ULK1) and to a lesser extent its paralog unc-51 like autophagy activating kinase 2 (ULK2). ULK1 is a serine/threonine kinase that is the core component of the initiation complex alongside autophagy related (ATG)13, Focal adhesion kinase family-interacting protein of 200 kDa (FIP200), and ATG101 [172,173,174,175]. During nutrient deprivation, autophagy is rapidly stimulated to allow for bulk degradation to supply provisions of amino acids for continued synthesis. Autophagy is negatively regulated by mTORC1 during nutritional and energetic surplus (Figure 3). However, amino acid insufficiency relieves mTORC1-mediated inhibitory phosphorylation of ULK1 and ATG13 [175], enhancing autophagy. Importantly, cellular energy status is also conveyed to the initiation complex through the energy sensor AMPK, which can inhibit mTORC1 or directly activate ULK1 via phosphorylation [176]. Apart from direct inhibitory phosphorylation of ULK1, mTORC1 also regulates ULK1 stability through inhibition of autophagy/Beclin 1 regulator (AMBRA1). AMBRA1 acts to facilitate TNF receptor associated factor 6 (TRAF6)-mediated ubiquitination of ULK1 and subsequent protein stabilization [177]. Upon ULK1 activation, ULK1 phosphorylates AMBRA1 and allows for the localization of the BECLIN 1-VPS34 core complex to autophagosome initiation sites primarily at the ER [178]. In addition to inducing the translocation of the BECLIN 1-VPS34 core complex, ULK1 also activates BECLIN 1 (Figure 3) and stimulates VPS34 class III phosphatidylinositol 3-kinase activity [179]. Production of phosphatidylinositol 3-phosphate at the site of phagophore initiation is a critical step that engages the WD-repeat protein interacting with phosphoinositides (WIPI) family of proteins which serve as effectors that recruit several ATG proteins and subsequently LC3 lipidation [180]. Lipidation of LC3 via conjugation to phosphatidylethanolamine plays a critical role in autophagosome formation through expansion of the isolation membrane [126,181,182,183].
While excessive autophagy can contribute to atrophy [151], in recent years selective autophagy of mitochondria, termed mitophagy, has gained attention as a necessary and critical quality control step for skeletal muscle health (Figure 3). Mitochondrial dysfunction that results in the loss of the mitochondrial membrane potential (Δψm) triggers a cascade of events that culminate in sequestration and autophagic clearance of damaged mitochondria. Under normal conditions, the serine/threonine kinase PTEN induced kinase 1 (PINK1) is sequestered within the mitochondrion [184] and its degradation depends on intact Δψm [185]. Upon mitochondrial damage and membrane depolarization, PINK1 is stabilized and accumulates in the outer mitochondrial membrane (OMM) [184,185,186] where it recruits the ubiquitin ligase Parkin (Figure 3). Parkin is normally cytosolic and inactive under control conditions [185]. Upon mitochondrial localization, Parkin is activated and decorates OMM proteins with polyubiquitin chains [187] (Figure 3), thereby promoting mitophagy through interactions with autophagy adapter proteins [188,189].
Investigations regarding whether skeletal muscle autophagy is reduced or enhanced with age have reached mixed conclusions [190,191,192,193,194,195]. While the consensus is that autophagy is blunted with aging, evidence for altered autophagy has primarily relied on the quantification of basal protein abundance of several autophagy effectors. However, changes to the abundance of autophagy regulators such as LC3-I/LC3-II do not accurately reflect autophagic flux within the tissue and instead represent a fixed snapshot in time of a highly dynamic process. Indeed, the LC3-I/LC3-II content may be enhanced due to greater LC3 lipidation or diminished lysosomal degradation. Autophagy can be more accurately estimated using autophagy inhibitors such as colchicine, allowing for measurements of accumulated autophagy effectors as an output for autophagy flux [196].

3.2.2. Defective Autophagy Impairs Skeletal Muscle Function and Mass

Measurements of autophagy flux have shown a trend for enhanced basal autophagy in aged muscle [105,193,197]. Nonetheless, it is intriguing that two effective interventions for mitigation of sarcopenia and aging, namely caloric restriction [198] and exercise [168,195,199,200], are known to stimulate skeletal muscle autophagy. Additionally, reducing autophagy in skeletal muscle recapitulates several aspects of sarcopenia. For example, autophagy deficient skeletal muscle of ATG7-knockout mice exhibit signs of myopathy alongside compromised skeletal muscle mass and function [201]. Skeletal muscle specific knockout of ATG7 in aged mice similarly reduces muscle mass and produces signs of neuromuscular junction dysfunction [195].
Skeletal muscle AMPK-knockout mice also have defective autophagy and exhibit age-related muscle dysfunction associated with signs of fiber degeneration, diminished force production, and greater mitochondrial size and Parkin accumulation at 18 months of age [202]. These data imply that relative to age matched wild types, reduced autophagy in skeletal muscle of AMPK-knockout mice results in impaired muscle function and mitochondrial clearance, which manifests as myopathy in an age-dependent manner [202]. Additionally, evidence from the literature indicates that enhancing autophagy ameliorates the aging phenotype. While activation of AMPK is known to inhibit mTORC1 and protein synthesis, it is interesting to note that muscle atrophy in obese sarcopenic rats was ameliorated by resveratrol, which was also found to prevent myotube atrophy in vitro partially via the PKA/LKB1/AMPK pathway [203].
Skeletal muscle expression of FoxO in drosophila enhances autophagy and delays the aging associated accumulation of protein aggregates in skeletal muscle [204]. In contrast, skeletal muscle FoxO1/3/4 triple-knockout mice exhibit spared muscle mass and preserved specific force, a functional measure of muscle quality, during aging without apparent signs of muscle degeneration, despite dramatically greater fiber central nucleation [205]. While autophagy flux was not compared in aged control and triple-knockout mice, young triple-knockout mice did not present an apparent defect in muscle autophagy. Thus, the preservation of skeletal muscle mass and function in FoxO triple-knockout mice may be due to an observed conservation of mitochondrial function without apparent changes to the autophagy machinery.

3.2.3. mTORC1 and Autophagy

One potential mechanism by which autophagy may be impaired in aged skeletal muscle is via sustained activation of mTORC1. Comparisons of autophagy and mTORC1 signaling in skeletal muscle of mice that are freely fed revealed little difference between young and old animals. However, upon fasting, skeletal muscle from young mice displayed diminished phosphorylation of AKT and the mTORC1 targets p70S6K1, ribosomal protein S6 (RPS6), and 4E-BP1. Notably, fasting-induced changes were absent in aged mice. Furthermore, mTORC1 phosphorylation of ULK1 (Ser757) was reduced by fasting in young but not aged mice [116]. Mirroring the sustained mTORC1 activity and reduced autophagic signaling that has been reported in aged animals, skeletal muscle TSC1-knockout animals exhibit impaired autophagy that is concomitant with late onset myopathy, both of which are reversed by rapamycin treatment [206]. Indeed, the enhanced UPS activity present in skeletal muscle TSC1-knockout mice may be an attempt to compensate for impaired autophagic clearance [180]. A recent study by Crombie et al. [207] revealed a previously unknown aspect of mTORC1 regulation of autophagy in skeletal muscle. Interestingly, activation of 4E-BP1 but not inhibition of p70S6K1 ameliorates the sarcopenic phenotype of skeletal muscle TSC1-knockout mice [207]. Despite reduced MPS with both p70S6K1 inhibition and 4E-BP1 activation, only activation of 4E-BP1 in TSC1-knockout muscle improved muscle size, function, and the induction of autophagy by fasting despite no change in ULK1 phosphorylation. TSC1 muscle-knockout mice exhibited greater protein aggregates and histological indications of accumulated mitochondria and lipofuscin; characteristic of blunted mitophagy and lysosome function. Importantly, activation of 4E-BP1 ameliorated signs of mitochondrial and lysosomal dysfunction present with mTORC1 hyperactivation, suggesting a critical role for 4E-BP1 modulation of autophagy that is independent of mTORC1-mediated ULK1 phosphorylation.

3.2.4. Autophagy and Mitochondrial Dysfunction

It is well established that mitochondrial dysfunction is a core characteristic of sarcopenia [126]. Whether aged skeletal muscle mitochondria exhibit a bioenergetic defect is debated, as respiration measured in isolated mitochondria produce a pronounced age associated defect that is less apparent in in situ preparations that maintain mitochondria in their native cellular environment [208]. This apparent discrepancy may be due to greater fragility of aged mitochondria that impairs their resilience to standard isolation procedures, thereby exacerbating respirometric defects [208,209]. Nonetheless, a greater susceptibility to apoptosis is a well-documented characteristic of mitochondria from aged muscle [126,210] which may be secondary to impaired mitophagy [210,211]. Additionally, while mitochondrial respiration is contingent on physical activity status in aged adults, sensitivity of mitochondria to permeability transition is enhanced by old age regardless of activity status [27], indicating that perturbations to mitochondrial health and apoptotic susceptibility exists in aged skeletal muscle independent of intrinsic bioenergetic capacity. While mitochondrial bioenergetic defects in aged skeletal muscle are not a universal finding, it is important to recognize that high-resolution respirometry, often used to measure mitochondrial function, commonly employs substrate-uncoupler-inhibitor-titration (SUIT) protocols that measure mitochondrial respiration in the presence of kinetically saturating concentrations of ADP that are several-fold higher than physiological ADP concentrations. Indeed, when tested at physiologically relevant ADP concentrations, in situ preparations of human aged skeletal muscle exhibit a respirometric defect and a higher Km for ADP when compared to young muscle [212], suggesting impaired ADP sensitivity in mitochondria from aged muscle. High-fat-diet feeding, which results in obesity and impairs mitochondrial health [213], also induces skeletal muscle respirometric defects at submaximal but not maximal ADP concentrations in rodents, despite the greater abundance of mitochondrial proteins [214]. Interestingly, in a murine model of sarcopenic obesity, mitochondrial uncoupling enhanced markers of mitophagy and improved skeletal muscle mass and function [215]. These data imply that the removal of dysfunctional mitochondria is critical for improving muscle function during sarcopenic obesity. Given that mitochondria exhibit a high spare capacity for ATP production during repeated muscle contraction, it is unclear if marginal reductions in bioenergetic function would perturb cellular homeostasis or produce an energetic defect at rest. However, the potential for greater reactive oxygen species production can induce damage to mitochondrial components for which there may be insufficient mitophagy to mitigate in aged muscle.

3.2.5. Ca2+ Dysregulation in Sarcopenia

Oxidative stress produced by the accumulation of dysfunctional mitochondria can also damage cellular structures that are near the damaged mitochondria. Indeed, in muscle, mitochondria share an intimate role with the ER/sarcoplasmic reticulum (SR) and influence Ca2+ homeostasis. Skeletal muscle function relies heavily on the propagation of signals from the neuromuscular junction to the transverse tubules to the SR, a specialized portion of the ER in striated muscle that mediates the release of Ca2+ ions through the ryanodine receptor to initiate muscle contraction. Disruption of Ca2+ handling by the SR can lead to elevated cytosolic Ca2+ and in turn, exacerbates mitochondrial reactive oxygen species (ROS) production. Oxidation of the ryanodine receptor is greater in aged than in young muscle, which can result in Ca2+ “leakage” from the SR into the cytoplasm [216]. Muscles from mice expressing a mutated form of the ryanodine receptor that enhances SR Ca2+ leakage present similar contractile deficits as aged mice [216]. In agreement with these findings, Delrio-Lorenzo et al. [217] found that skeletal muscle SR Ca2+ concentration decreases with age in drosophila melanogaster. Furthermore, the decrease of Ca2+ concentration was closely correlated with reduced physical function and not observed in neurons, highlighting a tissue specific dysregulation of Ca2+ homeostasis [217]. In addition to dysfunction at the ryanodine receptor, sarcopenia is also associated with reduced sarcoplasmic reticulum Ca2+ ATPase (SERCA) pump activity [218]. The SERCA pump is important for reuptake of Ca2+ from the cytosol to the SR at the end of an electrically evoked stimulus. Interestingly, treatment of aged mice with the allosteric SERCA activator CDN1163 improved indices of muscle function and attenuated mitochondrial dysfunction [218]. Furthermore, CuZnSOD deficient (Sod1-/-) mice, a model of premature aging, also exhibited reduced SERCA pump activity and skeletal muscle contractile deficits that are rescued by CDN1163 treatment [219]. Taken together, it appears that oxidative stress during advanced age perturbs SERCA pump function and elevates cytosolic skeletal muscle Ca2+, which can amplify mitochondrial dysfunction and contribute to declines in contractile function.
In addition to dysfunctional Ca2+ reuptake, there is evidence that store-operated Ca2+ entry (SOCE) is inefficient in sarcopenic muscle and may contribute to declines in skeletal muscle performance. SOCE is critical to maintain intracellular Ca2+ homeostasis as ER/SR Ca2+ stores are depleted. In brief, reductions in ER/SR Ca2+ are sensed by the Ca2+ sensor stromal interaction molecule 1 (STIM1), which resides on the ER/SR membrane and in response to Ca2+ depletion aggregates to ER/SR locales proximal to the plasma membrane [220]. STIM1 interacts with the plasma membrane channel Orai Ca2+ release-activated Ca2+ modulator 1 (Orai1) [221], which facilitates extracellular Ca2+ entry into the ER/SR. Indeed, the importance of SOCE for skeletal muscle health is apparent in mice with STIM1 haploinsufficiency, which exhibit markedly greater muscle fatigability [222]. Furthermore, mice that express a dominant negative form of Orai1 display both reduced skeletal muscle mass as well as enhanced susceptibility to fatigue during repeated muscle contraction [223]. Investigations regarding the relevance of SOCE to aging-associated functional deficits have revealed that skeletal muscle SOCE function is reduced in aged animals despite sustained STIM1 and Orai1 mRNA expression [224,225], although this has been challenged [226]. During ex vivo contractility assays, it was revealed that inhibition of SOCE reduces contractile activity in skeletal muscle from young but not aged animals, an effect that was most prominent at high frequency stimulation [225]. These data suggest that there exists a lack of SOCE contribution to contractile function at high intensities in geriatric muscle. Interestingly, aged skeletal muscle was shown to contain reduced abundance of the synaptophysin-related membrane protein Mitsugumin 29 (MG29) that regulates SR and transverse tubule contact sites [227]. A lack of MG29 has been shown to reduce skeletal muscle contractile function [228] as well as compromise SOCE [224]. These findings reveal that in addition to enhanced Ca2+ leakage from the ER/SR, inadequate SOCE may contribute to a dysregulation of Ca2+ homeostasis that compromises function in aged skeletal muscle.

4. Therapeutics for the Treatment of Sarcopenia

The evidence for non-pharmacological approaches to mitigating or preventing sarcopenia suggests that both exercise [229,230] and to a lesser extent caloric restriction [231,232,233] reduce age-associated muscle quality decline, although there exist concerns that a hypocaloric diet may also reduce lean mass [234,235]. Furthermore, although combined caloric restriction and exercise may yield the most benefit, attrition to lifestyle intervention may not be a practical solution for some individuals suffering from comorbidities or mobility impairments. Thus, pharmacological agents that mimic the benefits gained from routine exercise, caloric restriction, or dietary interventions are attractive therapeutic targets to prevent or mitigate sarcopenia (Figure 4).

4.1. Branched-Chain Amino Acid Supplementation

Branched-chain amino acids (BCAAs) play a crucial role as both signaling molecules and substrates for protein synthesis. Low blood levels of BCAAs are associated with sarcopenia and reduced physical function [236]. Ingestion of BCAAs was shown to enhance early post prandial myofibrillar protein synthesis in older males (~70 years old) [237]. Among the three BCAAs (leucine, valine, and isoleucine), leucine is considered to be the primary modulator of muscle anabolism. Analysis of dietary leucine intake and muscle mass have revealed that leucine intake is correlated to the preservation of lean mass over the course of 6 years in older (>65 years) but not younger subjects [238]. Additionally, Devries et al. [239] found that the acute induction of MPS in healthy older women (65–75 years) that ingested either 10 g of milk or 25 g of whey protein isolate was similar when leucine content was matched. These data reveal that the BCAA content of a meal is a primary determinant of the anabolic response to feeding. However, it is necessary to keep in mind that the consumption of BCAAs without dietary inclusion of other essential amino acids may not be as beneficial, as protein synthesis requires both anabolic signaling and adequate availability of other amino acids [240]. Nonetheless, short term supplementation of BCAAs in combination with other amino acids for 5 weeks revealed positive effects on sarcopenic parameters in elderly patients. However, the reported benefits were lost after 12 weeks of discontinuation [241]. Taken together, the available literature suggests that greater BCAA consumption, particularly that of leucine, in combination with adequate protein intake may be effective at preserving muscle mass in elderly individuals.

4.2. Rapamycin, Rapalogs, and Calorie Restriction

Recently, interest has grown over the use of mTORC1 inhibitors and calorie restriction mimetics as a potential treatment for sarcopenia. Indeed, the use of rapamycin and rapalogs have proven to be beneficial in ameliorating age-associated skeletal muscle defects in preclinical models [43,117,122]. Interestingly, while rapamycin is considered to act through a similar mechanism as calorie restriction, a recent study by Ham et al. [125] has revealed that rapamycin improves aging-associated skeletal muscle outcomes through mechanisms that are separate from caloric restriction. Rapamycin treatment tended to reverse age related gene expression patterns which were often augmented by caloric restriction in old animals, indicating divergent regulation of gene expression between the two interventions. Caloric restriction has been shown to reduce muscle fiber central nucleation and p62 protein accumulation in skeletal muscle TSC1-knockout mice, without altering Sqstm1 gene expression, autophagic flux, or diminishing mTORC1 activation [125]. These data indicate that caloric restriction can improve muscle degeneration through mechanisms that are independent of mTORC1 suppression. However, it is worth noting that improvements in skeletal muscle mass after caloric restriction were largely only positive when muscle mass was normalized to body mass, which tended to be significantly lower in aged calorie restricted mice. After caloric restriction, absolute mass of several hindlimb muscles in old mice were the same or diminished when compared to ad libitum fed old mice, which was also mirrored in fiber cross sectional area [125]. Furthermore, absolute force tended to be similar or lower in aged calorie restricted mice when compared to aged ad libitum mice, while specific force tended to improve [125]. These results suggest that while calorie restriction may negatively affect absolute muscle mass, relative muscle mass and quality may be improved during aging (Figure 4).

4.3. Mitochondrial Uncouplers

Indeed, mTORC1 inhibitors are not the only pharmaceutical agents with alleged benefits for aging associated muscle defects. Dietary supplementation with the mitochondrial uncoupler BAM15 induced markers of mitophagy and improved muscle mass in a murine model of sarcopenic obesity [215]. Administration of BAM15 reduced both ER stress and apoptotic signaling in the skeletal muscle of sarcopenic obese mice. Interestingly, while mTORC1 activation was not diminished there was a marked upregulation of AMPK phosphorylation concomitant with greater mitochondrial volume. Similar to BAM15, in vitro administration of urolithin A to the myoblast line of C2C12 cells reduces Δψm and enhances mitophagy. When given to aged mice on either a chow or a high-fat diet, dietary administration of urolithin A potently enhanced skeletal muscle physical function without changes to body mass. Improvements in physical function were concomitant with enhanced AMPK phosphorylation and greater abundance of markers of autophagy and mitophagy [242]. These data provide evidence that enhancing mitophagy to maintain the health of the skeletal muscle mitochondrial pool is advantageous for treatment of sarcopenic obesity (Figure 4).

4.4. Androgens and SARMs

Given the association between sarcopenia and hypogonadism, targeting of the androgen receptor via selective androgen receptor modulators (SARMs) and testosterone has been investigated as viable therapeutics for sarcopenia. The administration of exogenous testosterone, while known to substantially enhance skeletal muscle mass [243], is also associated with adverse outcomes regarding prostate health [244]. Co-administration of testosterone and the type II 5α-reductase inhibitor finasteride to older men (60–80 yrs. old) results in significant enhancements in skeletal muscle mass as well as strength without adverse outcomes in prostate size, indicating that androgen receptor activation in combination with type II 5α-reductase inhibitors may be a viable strategy to improve sarcopenia [245]. Activation of the androgen receptor using nonsteroidal SARMs has several benefits over testosterone due to their tissue selective anabolic properties. In female patients with sarcopenia, administration of the SARM MK-0773 for 6 months produced significant enhancements in lean body mass but failed to significantly improve physical function over placebo [246]. In contrast, the SARM GTx-024 improved both lean body mass and functional outcomes in healthy older men and women, which was concomitant with favorable metabolic indices (Figure 4).

4.5. Myostatin Inhibitors

Notably, there have also been several drug candidates produced to target myostatin, a potent negative regulator of skeletal muscle mass. Administration of the human monoclonal antibody REGN1033 to target myostatin in mice enhanced both muscle mass and force production and prevented muscle atrophy induced by limb immobilization. Treatment of aged mice with REGN1033 led to significantly increased muscle mass and force production without a loss in specific force, indicating a preservation of muscle quality [247]. Subcutaneous injections of the anti-myostatin antibody LY2495655 were tested in humans aged 75 and older who had a fall within the previous year and exhibited low muscle strength to determine how LY2495655 would alter appendicular lean mass and physical function [248]. The results indicated that LY2495655 treatment improved appendicular lean mass and physical performance measures as objective measures that heavily relied on power generation. Like antibodies that bind and neutralize myostatin, monoclonal antibodies that bind type II activin receptors and thus prevent myostatin associated signaling have also shown promise. For example, in a 24-week, randomized, double-blind, placebo-controlled study, the use of the type II activin receptor antagonist bimagrumab in sarcopenic community-living men and women aged 65 and older resulted in a greater accrual of lean body mass and appendicular lean mass than placebo [249]. The greater accrual of lean body mass and muscle volume were evident in as short as two weeks of treatment [249]. Additionally, while bimagrumab treatment did not improve six-minute walk distance between groups, there was a significant improvement in six-minute walk distance in the slowest walkers after bimagrumab treatment when compared to placebo controls. This suggests that bimagrumab treatment improved functional outcomes in those with the largest deficits in mobility based physical function. Additional work is needed to optimize bimagrumab, or other compounds that block type II activin receptors and evaluate muscle function, mobility, and muscle mass after long term treatment in the elderly. However, the initial data after acute administration suggest that this might be a beneficial approach for reducing sarcopenia (Figure 4).

4.6. Present Challenges and Future Direction

Currently, there are no approved drugs for the treatment of sarcopenia. Recent and past findings have shown that a finely tuned balance between anabolic and catabolic processes is required to maintain adequate muscle quality and mass in older ages. Questions remain regarding the appropriate degree of caloric restriction to combat age related muscle dysfunction. An important variable to consider when interpreting the impact of caloric restriction is to differentiate between the benefit of an energy deficit vs. the maintenance of a reduced body weight. A perpetual caloric deficit would result in a chronic loss of body mass and exacerbate sarcopenia. Thus, organisms adapt to reduced caloric intake via weight loss and energy sparing processes to reach an equilibrium between calorie expenditure and calorie intake. It is unclear how much of the purported benefit of calorie restriction is due to an incurred energy deficit vs. the maintenance of a healthier lower body weight (primarily reduced fat mass). Additionally, caloric restriction is difficult to execute consistently and effectively in a clinical setting. It is possible that regimens consisting of interspersed periods of caloric restriction with isocaloric diets may activate cellular pathways that confer protection against sarcopenia without compromising absolute skeletal muscle mass. Furthermore, while extensive lifestyle modification may not be a practical and viable option for elderly patients, small amounts of physical activity may improve autophagy/mitophagy and be augmented pharmaceutically to allow for even a greater benefit on muscle quality. A combination of exercise that is coupled with supplementation with the polyphenol resveratrol has also been shown to augment exercise adaptations in older men and women to offset sarcopenia [250]. This suggests that the benefit incurred from physical activity can be pharmaceutically amplified. Future research should elucidate interactions between pharmaceutical agents with exercise and/or dietary manipulation as potential avenues to combat sarcopenia.

5. Conclusions

The primary cause of cellular dysregulation that contributes to sarcopenia development remains elusive. However, recent findings implicate dysregulated proteostasis as an important characteristic of skeletal muscle in aged hosts that may contribute to muscle atrophy and functional decline leading to sarcopenia. Additionally, given the multi-factorial aspect of sarcopenia, effective treatment will likely be best achieved by addressing multiple pathways that contribute to the loss of muscle function and mass with aging. Furthering our understanding of the causes and consequence of dysfunctional proteostasis in aged skeletal muscle will allow for the development of targeted treatments to ameliorate one of the most debilitating geriatric conditions.

Author Contributions

Conceptualization, H.G.P. methodology, H.G.P. and S.E.A.; software, H.G.P.; validation, H.G.P. formal analysis, H.G.P.; investigation, H.G.P., C.R.P. and S.E.A.; resources, S.E.A. data curation, H.G.P. writing—original draft preparation, H.G.P.; writing—review and editing, H.G.P., C.R.P. and S.E.A.; visualization, H.G.P.; supervision, S.E.A.; project administration, S.E.A.; funding acquisition, S.E.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by W81XWH2110187 from the U.S. Department of Defense awarded to S.E.A.

Data Availability Statement

The data presented and reviewed in this manuscript are openly available in published forms as documented in the references of this manuscript.

Acknowledgments

The figures were created with BioRender.com.

Conflicts of Interest

The authors declare that there are no conflict of interest.

Abbreviations

mTOR, mammalian Target of Rapamycin; mTORC1, mammalian Target of Rapamycin Complex 1; MPS, muscle protein synthesis; MPB, muscle protein breakdown; PI3K, Phosphoinositide 3-kinase; PKB/Akt, Protein Kinase B; TSC1/2, Tuberous Sclerosis Complex 1/2; GAP, GTPase-Activating Protein; Rheb, Ras Homolog Enriched in Brain; GATOR 1/2, GTPase-activating protein activity toward Rags 1/2; CASTOR1, Cytosolic Arginine Sensor For mTORC1 Subunit 1; 4E-BP1, eukaryotic translation initiation factor 4E (eIF4E)-binding protein 1; p70S6K1, p70 ribosomal protein S6 kinase; TNFα, Tumor Necrosis Factor α; PGC, Peroxisome Proliferator-activated receptor-Gamma Coactivator; MuRF1, Muscle-specific RING Finger protein 1; ER, endoplasmic reticulum; RAPTOR, Regulatory Associated Protein of mTOR; UPS, ubiquitin proteasome system; MyoD1, Myogenic Differentiation 1; eIF3-f, eukaryotic translation initiation factor 3 subunit F; FoxO, Forkhead box O; AMPK, AMP-activated Protein Kinase; NRF1, Nuclear Factor Erythroid 2-Like 1; ULK1/2, unc-51 like autophagy activating kinase 1/2; ATG, Autophagy related; AMBRA1, autophagy/beclin 1 regulator; TRAF6, TNF receptor associated factor 6; WIPI, WD-repeat Protein Interacting with Phosphoinositides; PINK1, PTEN induced kinase 1; OMM, outer mitochondrial membrane; RPS6, Ribosomal Protein S6; SUIT, substrate-uncoupler-inhibitor-titration; AR, Androgen Receptor; MSTN, Myostatin; SARM, Selective Androgen Receptor Modulator; DHT, Dihydrotestosterone; IRS-1, Insulin receptor substrate 1; SR, sarcoplasmic reticulum; SERCA, sarcoplasmic reticulum Ca2+ ATPase; ROS, reactive oxygen species: BCAA, branched-chain amino acid; SOCE, store-operated Ca2+ entry; MG29, Mitsugumin 29.

References

  1. Janssen, I.; Heymsfield, S.B.; Wang, Z.; Ross, R. Skeletal muscle mass and distribution in 468 men and women aged 18–88 yr. J. Appl. Physiol. 2000, 89, 81–88. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. DeFronzo, R.A.; Tripathy, D. Skeletal muscle insulin resistance is the primary defect in type 2 diabetes. Diabetes Care 2009, 32, S157–S163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Li, R.; Xia, J.; Zhang, X.; Gathirua-Mwangi, W.G.; Guo, J.; Li, Y.; McKenzie, S.; Song, Y. Associations of muscle mass and strength with all-cause mortality among US older adults. Med. Sci. Sport. Exerc. 2018, 50, 458–467. [Google Scholar] [CrossRef] [PubMed]
  4. Szulc, P.; Munoz, F.; Marchand, F.; Chapurlat, R.; Delmas, P.D. Rapid loss of appendicular skeletal muscle mass is associated with higher all-cause mortality in older men: The prospective MINOS study. Am. J. Clin. Nutr. 2010, 91, 1227–1236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Landi, F.; Liperoti, R.; Russo, A.; Giovannini, S.; Tosato, M.; Capoluongo, E.; Bernabei, R.; Onder, G. Sarcopenia as a risk factor for falls in elderly individuals: Results from the ilSIRENTE study. Clin. Nutr. 2012, 31, 652–658. [Google Scholar] [CrossRef]
  6. Koo, B.K.; Roh, E.; Yang, Y.S.; Moon, M.K. Difference between old and young adults in contribution of β-cell function and sarcopenia in developing diabetes mellitus. J. Diabetes Investig. 2016, 7, 233–240. [Google Scholar] [CrossRef]
  7. Cruz-Jentoft, A.J.; Bahat, G.; Bauer, J.; Boirie, Y.; Bruyère, O.; Cederholm, T.; Cooper, C.; Landi, F.; Rolland, Y.; Sayer, A.A. Sarcopenia: Revised European consensus on definition and diagnosis. Age Ageing 2019, 48, 16–31. [Google Scholar] [CrossRef] [Green Version]
  8. Koster, A.; Ding, J.; Stenholm, S.; Caserotti, P.; Houston, D.K.; Nicklas, B.J.; You, T.; Lee, J.S.; Visser, M.; Newman, A.B. Does the amount of fat mass predict age-related loss of lean mass, muscle strength, and muscle quality in older adults? J. Gerontol. Ser. A Biomed. Sci. Med. Sci. 2011, 66, 888–895. [Google Scholar] [CrossRef]
  9. Delmonico, M.J.; Harris, T.B.; Visser, M.; Park, S.W.; Conroy, M.B.; Velasquez-Mieyer, P.; Boudreau, R.; Manini, T.M.; Nevitt, M.; Newman, A.B.; et al. Longitudinal study of muscle strength, quality, and adipose tissue infiltration. Am. J. Clin. Nutr. 2009, 90, 1579–1585. [Google Scholar]
  10. Mitchell, W.K.; Williams, J.; Atherton, P.; Larvin, M.; Lund, J.; Narici, M. Sarcopenia, dynapenia, and the impact of advancing age on human skeletal muscle size and strength; a quantitative review. Front. Physiol. 2012, 3, 260. [Google Scholar] [CrossRef] [Green Version]
  11. McCastlain, K.; Howell, C.R.; Welsh, C.E.; Wang, Z.; Wilson, C.L.; Mulder, H.L.; Easton, J.; Mertens, A.C.; Zhang, J.; Yasui, Y. The association of mitochondrial copy number with sarcopenia in adult survivors of childhood cancer. JNCI J. Natl. Cancer Inst. 2021, 113, 1570–1580. [Google Scholar] [CrossRef]
  12. Ooi, P.H.; Thompson-Hodgetts, S.; Pritchard-Wiart, L.; Gilmour, S.M.; Mager, D.R. Pediatric sarcopenia: A paradigm in the overall definition of malnutrition in children? J. Parenter. Enter. Nutr. 2020, 44, 407–418. [Google Scholar] [CrossRef]
  13. Suzuki, D.; Kobayashi, R.; Sano, H.; Hori, D.; Kobayashi, K. Sarcopenia after induction therapy in childhood acute lymphoblastic leukemia: Its clinical significance. Int. J. Hematol. 2018, 107, 486–489. [Google Scholar] [CrossRef]
  14. Bischoff-Ferrari, H.A.; Orav, J.; Kanis, J.A.; Rizzoli, R.; Schlögl, M.; Staehelin, H.; Willett, W.C.; Dawson-Hughes, B. Comparative performance of current definitions of sarcopenia against the prospective incidence of falls among community-dwelling seniors age 65 and older. Osteoporos. Int. 2015, 26, 2793–2802. [Google Scholar] [CrossRef]
  15. Schaap, L.A.; Van Schoor, N.M.; Lips, P.; Visser, M. Associations of sarcopenia definitions, and their components, with the incidence of recurrent falling and fractures: The longitudinal aging study Amsterdam. J. Gerontol. Ser. A 2018, 73, 1199–1204. [Google Scholar] [CrossRef]
  16. Morley, J.E.; Abbatecola, A.M.; Argiles, J.M.; Baracos, V.; Bauer, J.; Bhasin, S.; Cederholm, T.; Coats, A.J.S.; Cummings, S.R.; Evans, W.J. Sarcopenia with limited mobility: An international consensus. J. Am. Med. Dir. Assoc. 2011, 12, 403–409. [Google Scholar] [CrossRef] [Green Version]
  17. Lauretani, F.; Russo, C.R.; Bandinelli, S.; Bartali, B.; Cavazzini, C.; Di Iorio, A.; Corsi, A.M.; Rantanen, T.; Guralnik, J.M.; Ferrucci, L. Age-associated changes in skeletal muscles and their effect on mobility: An operational diagnosis of sarcopenia. J. Appl. Physiol. 2003, 95, 1851–1860. [Google Scholar] [CrossRef] [Green Version]
  18. Janssen, I.; Heymsfield, S.B.; Ross, R. Low relative skeletal muscle mass (sarcopenia) in older persons is associated with functional impairment and physical disability. J. Am. Geriatr. Soc. 2002, 50, 889–896. [Google Scholar] [CrossRef] [Green Version]
  19. 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]
  20. Goates, S.; Du, K.; Arensberg, M.; Gaillard, T.; Guralnik, J.; Pereira, S.L. Economic impact of hospitalizations in US adults with sarcopenia. J. Frailty Aging 2019, 8, 93–99. [Google Scholar] [CrossRef]
  21. Janssen, I.; Shepard, D.S.; Katzmarzyk, P.T.; Roubenoff, R. The healthcare costs of sarcopenia in the United States. J. Am. Geriatr. Soc. 2004, 52, 80–85. [Google Scholar] [CrossRef] [PubMed]
  22. World Health Organization. Ageing and Health; World Health Organization: Geneva, Switzerland, 2018.
  23. Idoate, F.; Cadore, E.L.; Casas-Herrero, A.; Zambom-Ferraresi, F.; Marcellán, T.; de Gordoa, A.R.; Rodriguez-Mañas, L.; Bastarrika, G.; Marques, M.C.; Martínez-Velilla, N. Adipose tissue compartments, muscle mass, muscle fat infiltration, and coronary calcium in institutionalized frail nonagenarians. Eur. Radiol. 2015, 25, 2163–2175. [Google Scholar] [CrossRef] [PubMed]
  24. Landi, F.; Calvani, R.; Cesari, M.; Tosato, M.; Martone, A.M.; Bernabei, R.; Onder, G.; Marzetti, E. Sarcopenia as the biological substrate of physical frailty. Clin. Geriatr. Med. 2015, 31, 367–374. [Google Scholar] [CrossRef] [PubMed]
  25. Deschenes, M.R.; Roby, M.A.; Eason, M.K.; Harris, M.B. Remodeling of the neuromuscular junction precedes sarcopenia related alterations in myofibers. Exp. Gerontol. 2010, 45, 389–393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Tamaki, T.; Hirata, M.; Uchiyama, Y. Qualitative alteration of peripheral motor system begins prior to appearance of typical sarcopenia syndrome in middle-aged rats. Front. Aging Neurosci. 2014, 6, 296. [Google Scholar] [CrossRef]
  27. Spendiff, S.; Vuda, M.; Gouspillou, G.; Aare, S.; Perez, A.; Morais, J.A.; Jagoe, R.T.; Filion, M.E.; Glicksman, R.; Kapchinsky, S. Denervation drives mitochondrial dysfunction in skeletal muscle of octogenarians. J. Physiol. 2016, 594, 7361–7379. [Google Scholar] [CrossRef] [Green Version]
  28. Visser, M.; Pahor, M.; Taaffe, D.R.; Goodpaster, B.H.; Simonsick, E.M.; Newman, A.B.; Nevitt, M.; Harris, T.B. Relationship of interleukin-6 and tumor necrosis factor-α with muscle mass and muscle strength in elderly men and women: The Health ABC Study. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 2002, 57, M326–M332. [Google Scholar] [CrossRef] [Green Version]
  29. Li, C.W.; Yu, K.; Shyh-Chang, N.; Li, G.X.; Jiang, L.J.; Yu, S.L.; Xu, L.Y.; Liu, R.J.; Guo, Z.J.; Xie, H.Y. Circulating factors associated with sarcopenia during ageing and after intensive lifestyle intervention. J. Cachexia Sarcopenia Muscle 2019, 10, 586–600. [Google Scholar] [CrossRef] [Green Version]
  30. Dalle, S.; Rossmeislova, L.; Koppo, K. The role of inflammation in age-related sarcopenia. Front. Physiol. 2017, 8, 1045. [Google Scholar] [CrossRef] [Green Version]
  31. Pistilli, E.E.; Jackson, J.R.; Alway, S.E. Death receptor-associated pro-apoptotic signaling in aged skeletal muscle. Apoptosis 2006, 11, 2115–2126. [Google Scholar] [CrossRef]
  32. Pistilli, E.E.; Siu, P.M.; Alway, S.E. Molecular regulation of apoptosis in fast plantaris muscles of aged rats. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 2006, 61, 245–255. [Google Scholar] [CrossRef]
  33. Alway, S.E.; Siu, P.M. Nuclear apoptosis contributes to sarcopenia. Exerc. Sport Sci. Rev. 2008, 36, 51–57. [Google Scholar] [CrossRef] [Green Version]
  34. Bian, A.; Ma, Y.; Zhou, X.; Guo, Y.; Wang, W.; Zhang, Y.; Wang, X. Association between sarcopenia and levels of growth hormone and insulin-like growth factor-1 in the elderly. BMC Musculoskelet. Disord. 2020, 21, 1–9. [Google Scholar] [CrossRef] [Green Version]
  35. Yuki, A.; Otsuka, R.; Kozakai, R.; Kitamura, I.; Okura, T.; Ando, F.; Shimokata, H. Relationship between low free testosterone levels and loss of muscle mass. Sci. Rep. 2013, 3, 1–5. [Google Scholar] [CrossRef] [Green Version]
  36. Kong, S.H.; Kim, J.H.; Lee, J.H.; Hong, A.R.; Shin, C.S.; Cho, N.H. Dehydroepiandrosterone sulfate and free testosterone but not estradiol are related to muscle strength and bone microarchitecture in older adults. Calcif. Tissue Int. 2019, 105, 285–293. [Google Scholar] [CrossRef]
  37. Gianoudis, J.; Bailey, C.; Daly, R. Associations between sedentary behaviour and body composition, muscle function and sarcopenia in community-dwelling older adults. Osteoporos. Int. 2015, 26, 571–579. [Google Scholar] [CrossRef]
  38. Distefano, G.; Standley, R.A.; Zhang, X.; Carnero, E.A.; Yi, F.; Cornnell, H.H.; Coen, P.M. Physical activity unveils the relationship between mitochondrial energetics, muscle quality, and physical function in older adults. J. Cachexia Sarcopenia Muscle 2018, 9, 279–294. [Google Scholar] [CrossRef] [Green Version]
  39. Landi, F.; Liperoti, R.; Russo, A.; Giovannini, S.; Tosato, M.; Barillaro, C.; Capoluongo, E.; Bernabei, R.; Onder, G. Association of anorexia with sarcopenia in a community-dwelling elderly population: Results from the ilSIRENTE study. Eur. J. Nutr. 2013, 52, 1261–1268. [Google Scholar] [CrossRef]
  40. Morley, J.E. Anorexia of ageing: A key component in the pathogenesis of both sarcopenia and cachexia. J. Cachexia Sarcopenia Muscle 2017, 8, 523–526. [Google Scholar] [CrossRef] [Green Version]
  41. Coelho, H.J., Jr.; Calvani, R.; Azzolino, D.; Picca, A.; Tosato, M.; Landi, F.; Cesari, M.; Marzetti, E. Protein Intake and Sarcopenia in Older Adults: A Systematic Review and Meta-Analysis. Int. J. Environ. Res. Public Health 2022, 19, 8718. [Google Scholar] [CrossRef]
  42. Drummond, M.J.; Fry, C.S.; Glynn, E.L.; Dreyer, H.C.; Dhanani, S.; Timmerman, K.L.; Volpi, E.; Rasmussen, B.B. Rapamycin administration in humans blocks the contraction-induced increase in skeletal muscle protein synthesis. J. Physiol. 2009, 587, 1535–1546. [Google Scholar] [CrossRef] [PubMed]
  43. Ham, D.J.; Börsch, A.; Lin, S.; Thürkauf, M.; Weihrauch, M.; Reinhard, J.R.; Delezie, J.; Battilana, F.; Wang, X.; Kaiser, M.S. The neuromuscular junction is a focal point of mTORC1 signaling in sarcopenia. Nat. Commun. 2020, 11, 4510. [Google Scholar] [CrossRef] [PubMed]
  44. Manning, B.D.; Tee, A.R.; Logsdon, M.N.; Blenis, J.; Cantley, L.C. Identification of the tuberous sclerosis complex-2 tumor suppressor gene product tuberin as a target of the phosphoinositide 3-kinase/akt pathway. Mol. Cell 2002, 10, 151–162. [Google Scholar] [CrossRef] [PubMed]
  45. Saucedo, L.J.; Gao, X.; Chiarelli, D.A.; Li, L.; Pan, D.; Edgar, B.A. Rheb promotes cell growth as a component of the insulin/TOR signalling network. Nat. Cell Biol. 2003, 5, 566–571. [Google Scholar] [CrossRef] [PubMed]
  46. Bar-Peled, L.; Chantranupong, L.; Cherniack, A.D.; Chen, W.W.; Ottina, K.A.; Grabiner, B.C.; Spear, E.D.; Carter, S.L.; Meyerson, M.; Sabatini, D.M. A Tumor suppressor complex with GAP activity for the Rag GTPases that signal amino acid sufficiency to mTORC1. Science 2013, 340, 1100–1106. [Google Scholar] [CrossRef] [Green Version]
  47. Wolfson, R.L.; Chantranupong, L.; Saxton, R.A.; Shen, K.; Scaria, S.M.; Cantor, J.R.; Sabatini, D.M. Sestrin2 is a leucine sensor for the mTORC1 pathway. Science 2016, 351, 43–48. [Google Scholar] [CrossRef] [Green Version]
  48. Saxton, R.A.; Chantranupong, L.; Knockenhauer, K.E.; Schwartz, T.U.; Sabatini, D.M. Mechanism of arginine sensing by CASTOR1 upstream of mTORC1. Nature 2016, 536, 229–233. [Google Scholar] [CrossRef] [Green Version]
  49. Chantranupong, L.; Scaria, S.M.; Saxton, R.A.; Gygi, M.P.; Shen, K.; Wyant, G.A.; Wang, T.; Harper, J.W.; Gygi, S.P.; Sabatini, D.M. The CASTOR proteins are arginine sensors for the mTORC1 pathway. Cell 2016, 165, 153–164. [Google Scholar] [CrossRef] [Green Version]
  50. Gingras, A.-C.; Raught, B.; Gygi, S.P.; Niedzwiecka, A.; Miron, M.; Burley, S.K.; Polakiewicz, R.D.; Wyslouch-Cieszynska, A.; Aebersold, R.; Sonenberg, N. Hierarchical phosphorylation of the translation inhibitor 4E-BP1. Genes Dev. 2001, 15, 2852–2864. [Google Scholar] [CrossRef]
  51. Browne, G.J.; Proud, C.G. A novel mTOR-regulated phosphorylation site in elongation factor 2 kinase modulates the activity of the kinase and its binding to calmodulin. Mol. Cell. Biol. 2004, 24, 2986–2997. [Google Scholar] [CrossRef] [Green Version]
  52. Chauvin, C.; Koka, V.; Nouschi, A.; Mieulet, V.; Hoareau-Aveilla, C.; Dreazen, A.; Cagnard, N.; Carpentier, W.; Kiss, T.; Meyuhas, O. Ribosomal protein S6 kinase activity controls the ribosome biogenesis transcriptional program. Oncogene 2014, 33, 474–483. [Google Scholar] [CrossRef]
  53. Holz, M.K.; Ballif, B.A.; Gygi, S.P.; Blenis, J. mTOR and S6K1 mediate assembly of the translation preinitiation complex through dynamic protein interchange and ordered phosphorylation events. Cell 2005, 123, 569–580. [Google Scholar] [CrossRef] [Green Version]
  54. Wang, X.; Li, W.; Williams, M.; Terada, N.; Alessi, D.R.; Proud, C.G. Regulation of elongation factor 2 kinase by p90(RSK1) and p70 S6 kinase. EMBO J. 2001, 20, 4370–4379. [Google Scholar] [CrossRef]
  55. Mosoni, L.; Valluy, M.; Serrurier, B.; Prugnaud, J.; Obled, C.; Guezennec, C.-Y.; Mirand, P.P. Altered response of protein synthesis to nutritional state and endurance training in old rats. Am. J. Physiol.-Endocrinol. Metab. 1995, 268, E328–E335. [Google Scholar] [CrossRef]
  56. Volpi, E.; Mittendorfer, B.; Rasmussen, B.B.; Wolfe, R.R. The response of muscle protein anabolism to combined hyperaminoacidemia and glucose-induced hyperinsulinemia is impaired in the elderly. J. Clin. Endocrinol. Metab. 2000, 85, 4481–4490. [Google Scholar] [CrossRef] [Green Version]
  57. Cuthbertson, D.; Smith, K.; Babraj, J.; Leese, G.; Waddell, T.; Atherton, P.; Wackerhage, H.; Taylor, P.M.; Rennie, M.J. Anabolic signaling deficits underlie amino acid resistance of wasting, aging muscle. FASEB J. 2005, 19, 422–424. [Google Scholar] [CrossRef] [Green Version]
  58. Volpi, E.; Sheffield-Moore, M.; Rasmussen, B.B.; Wolfe, R.R. Basal muscle amino acid kinetics and protein synthesis in healthy young and older men. JAMA 2001, 286, 1206–1212. [Google Scholar] [CrossRef]
  59. Moore, D.R.; Churchward-Venne, T.A.; Witard, O.; Breen, L.; Burd, N.A.; Tipton, K.D.; Phillips, S.M. Protein ingestion to stimulate myofibrillar protein synthesis requires greater relative protein intakes in healthy older versus younger men. J. Gerontol. Ser. A Biomed. Sci. Med. Sci. 2015, 70, 57–62. [Google Scholar] [CrossRef] [Green Version]
  60. Katsanos, C.S.; Kobayashi, H.; Sheffield-Moore, M.; Aarsland, A.; Wolfe, R.R. Aging is associated with diminished accretion of muscle proteins after the ingestion of a small bolus of essential amino acids. Am. J. Clin. Nutr. 2005, 82, 1065–1073. [Google Scholar] [CrossRef] [Green Version]
  61. Szwiega, S.; Pencharz, P.B.; Rafii, M.; Lebarron, M.; Chang, J.; Ball, R.O.; Kong, D.; Xu, L.; Elango, R.; Courtney-Martin, G. Dietary leucine requirement of older men and women is higher than current recommendations. Am. J. Clin. Nutr. 2021, 113, 410–419. [Google Scholar] [CrossRef]
  62. Evenson, K.R.; Buchner, D.M.; Morland, K.B. Objective measurement of physical activity and sedentary behavior among US adults aged 60 years or older. Prev. Chronic Dis. 2012, 9, E26. [Google Scholar] [CrossRef] [PubMed]
  63. Breen, L.; Stokes, K.A.; Churchward-Venne, T.A.; Moore, D.R.; Baker, S.K.; Smith, K.; Atherton, P.J.; Phillips, S.M. Two weeks of reduced activity decreases leg lean mass and induces “anabolic resistance” of myofibrillar protein synthesis in healthy elderly. J. Clin. Endocrinol. Metab. 2013, 98, 2604–2612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Dardevet, D.; Sornet, C.; Balage, M.; Grizard, J. Stimulation of in vitro rat muscle protein synthesis by leucine decreases with age. J. Nutr. 2000, 130, 2630–2635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Katsanos, C.S.; Kobayashi, H.; Sheffield-Moore, M.; Aarsland, A.; Wolfe, R.R. A high proportion of leucine is required for optimal stimulation of the rate of muscle protein synthesis by essential amino acids in the elderly. Am. J. Physiol.-Endocrinol. Metab. 2006, 291, E381–E387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Moro, T.; Brightwell, C.R.; Deer, R.R.; Graber, T.G.; Galvan, E.; Fry, C.S.; Volpi, E.; Rasmussen, B.B. Muscle protein anabolic resistance to essential amino acids does not occur in healthy older adults before or after resistance exercise training. J. Nutr. 2018, 148, 900–909. [Google Scholar] [CrossRef] [Green Version]
  67. Kim, T.N.; Park, M.S.; Lim, K.I.; Choi, H.Y.; Yang, S.J.; Yoo, H.J.; Kang, H.J.; Song, W.; Choi, H.; Baik, S.H. Relationships between sarcopenic obesity and insulin resistance, inflammation, and vitamin D status: The K orean S arcopenic O besity S tudy. Clin. Endocrinol. 2013, 78, 525–532. [Google Scholar] [CrossRef]
  68. Petersen, K.F.; Befroy, D.; Dufour, S.; Dziura, J.; Ariyan, C.; Rothman, D.L.; DiPietro, L.; Cline, G.W.; Shulman, G.I. Mitochondrial dysfunction in the elderly: Possible role in insulin resistance. Science 2003, 300, 1140–1142. [Google Scholar] [CrossRef] [Green Version]
  69. Vieira-Lara, M.A.; Dommerholt, M.B.; Zhang, W.; Blankestijn, M.; Wolters, J.C.; Abegaz, F.; Gerding, A.; van der Veen, Y.T.; Thomas, R.; van Os, R.P. Age-related susceptibility to insulin resistance arises from a combination of CPT1B decline and lipid overload. BMC Biol. 2021, 19, 154. [Google Scholar] [CrossRef]
  70. Messa, G.A.; Piasecki, M.; Hurst, J.; Hill, C.; Tallis, J.; Degens, H. The impact of a high-fat diet in mice is dependent on duration and age, and differs between muscles. J. Exp. Biol. 2020, 223, jeb217117. [Google Scholar] [CrossRef] [Green Version]
  71. Stephens, F.B.; Chee, C.; Wall, B.T.; Murton, A.J.; Shannon, C.E.; Van Loon, L.J.; Tsintzas, K. Lipid-induced insulin resistance is associated with an impaired skeletal muscle protein synthetic response to amino acid ingestion in healthy young men. Diabetes 2015, 64, 1615–1620. [Google Scholar] [CrossRef] [Green Version]
  72. Beals, J.W.; Sukiennik, R.A.; Nallabelli, J.; Emmons, R.S.; Van Vliet, S.; Young, J.R.; Ulanov, A.V.; Li, Z.; Paluska, S.A.; De Lisio, M. Anabolic sensitivity of postprandial muscle protein synthesis to the ingestion of a protein-dense food is reduced in overweight and obese young adults. Am. J. Clin. Nutr. 2016, 104, 1014–1022. [Google Scholar] [CrossRef] [Green Version]
  73. Murton, A.J.; Marimuthu, K.; Mallinson, J.E.; Selby, A.L.; Smith, K.; Rennie, M.J.; Greenhaff, P.L. Obesity appears to be associated with altered muscle protein synthetic and breakdown responses to increased nutrient delivery in older men, but not reduced muscle mass or contractile function. Diabetes 2015, 64, 3160–3171. [Google Scholar] [CrossRef] [Green Version]
  74. Smeuninx, B.; Mckendry, J.; Wilson, D.; Martin, U.; Breen, L. Age-related anabolic resistance of myofibrillar protein synthesis is exacerbated in obese inactive individuals. J. Clin. Endocrinol. Metab. 2017, 102, 3535–3545. [Google Scholar] [CrossRef] [Green Version]
  75. Purves-Smith, F.M.; Solbak, N.M.; Rowan, S.L.; Hepple, R.T. Severe atrophy of slow myofibers in aging muscle is concealed by myosin heavy chain co-expression. Exp. Gerontol. 2012, 47, 913–918. [Google Scholar] [CrossRef]
  76. Siddharth, J.; Chakrabarti, A.; Pannerec, A.; Karaz, S.; Morin-Rivron, D.; Masoodi, M.; Feige, J.N.; Parkinson, S.J. Aging and sarcopenia associate with specific interactions between gut microbes, serum biomarkers and host physiology in rats. Aging 2017, 9, 1698–1720. [Google Scholar] [CrossRef] [Green Version]
  77. Ni, Y.; Yang, X.; Zheng, L.; Wang, Z.; Wu, L.; Jiang, J.; Yang, T.; Ma, L.; Fu, Z. Lactobacillus and Bifidobacterium improves physiological function and cognitive ability in aged mice by the regulation of gut microbiota. Mol. Nutr. Food Res. 2019, 63, 1900603. [Google Scholar] [CrossRef]
  78. Lahiri, S.; Kim, H.; Garcia-Perez, I.; Reza, M.M.; Martin, K.A.; Kundu, P.; Cox, L.M.; Selkrig, J.; Posma, J.M.; Zhang, H. The gut microbiota influences skeletal muscle mass and function in mice. Sci. Transl. Med. 2019, 11, eaan5662. [Google Scholar] [CrossRef] [Green Version]
  79. Huang, W.-C.; Chen, Y.-H.; Chuang, H.-L.; Chiu, C.-C.; Huang, C.-C. Investigation of the effects of microbiota on exercise physiological adaption, performance, and energy utilization using a gnotobiotic animal model. Front. Microbiol. 2019, 10, 1906. [Google Scholar] [CrossRef] [Green Version]
  80. Stanley, D.; Denman, S.E.; Hughes, R.J.; Geier, M.S.; Crowley, T.M.; Chen, H.; Haring, V.R.; Moore, R.J. Intestinal microbiota associated with differential feed conversion efficiency in chickens. Appl. Microbiol. Biotechnol. 2012, 96, 1361–1369. [Google Scholar] [CrossRef]
  81. Thevaranjan, N.; Puchta, A.; Schulz, C.; Naidoo, A.; Szamosi, J.; Verschoor, C.P.; Loukov, D.; Schenck, L.P.; Jury, J.; Foley, K.P. Age-associated microbial dysbiosis promotes intestinal permeability, systemic inflammation, and macrophage dysfunction. Cell Host Microbe 2017, 21, 455–466.e4. [Google Scholar] [CrossRef] [Green Version]
  82. Sovran, B.; Hugenholtz, F.; Elderman, M.; Van Beek, A.A.; Graversen, K.; Huijskes, M.; Boekschoten, M.V.; Savelkoul, H.F.; De Vos, P.; Dekker, J. Age-associated impairment of the mucus barrier function is associated with profound changes in microbiota and immunity. Sci. Rep. 2019, 9, 1437. [Google Scholar] [CrossRef] [PubMed]
  83. Ghosh, S.; Lertwattanarak, R.; Garduño, J.d.J.; Galeana, J.J.; Li, J.; Zamarripa, F.; Lancaster, J.L.; Mohan, S.; Hussey, S.; Musi, N. Elevated muscle TLR4 expression and metabolic endotoxemia in human aging. J. Gerontol. Ser. A Biomed. Sci. Med. Sci. 2015, 70, 232–246. [Google Scholar] [CrossRef] [PubMed]
  84. Biagi, E.; Nylund, L.; Candela, M.; Ostan, R.; Bucci, L.; Pini, E.; Nikkïla, J.; Monti, D.; Satokari, R.; Franceschi, C. Through ageing, and beyond: Gut microbiota and inflammatory status in seniors and centenarians. PLoS ONE 2010, 5, e10667. [Google Scholar] [CrossRef]
  85. Claesson, M.J.; Jeffery, I.B.; Conde, S.; Power, S.E.; O’connor, E.M.; Cusack, S.; Harris, H.; Coakley, M.; Lakshminarayanan, B.; O’sullivan, O. Gut microbiota composition correlates with diet and health in the elderly. Nature 2012, 488, 178–184. [Google Scholar] [CrossRef] [PubMed]
  86. Kilroe, S.P.; Fulford, J.; Jackman, S.R.; Van Loon, L.J.; Wall, B.T. Temporal muscle-specific disuse atrophy during one week of leg immobilization. Med. Sci. Sport. Exerc. 2020, 52, 944–954. [Google Scholar] [CrossRef]
  87. Lang, F.; Aravamudhan, S.; Nolte, H.; Tuerk, C.; Hölper, S.; Müller, S.; Günther, S.; Blaauw, B.; Braun, T.; Krüger, M. Dynamic changes in the mouse skeletal muscle proteome during denervation-induced atrophy. Dis. Model. Mech. 2017, 10, 881–896. [Google Scholar]
  88. Siu, P.M.; Alway, S.E. Mitochondria-associated apoptotic signalling in denervated rat skeletal muscle. J. Physiol. 2005, 565, 309–323. [Google Scholar] [CrossRef]
  89. Chen, K.D.; Alway, S.E. Clenbuterol reduces soleus muscle fatigue during disuse in aged rats. Muscle Nerve Off. J. Am. Assoc. Electrodiagn. Med. 2001, 24, 211–222. [Google Scholar] [CrossRef]
  90. Myers, M.J.; Shepherd, D.L.; Durr, A.J.; Stanton, D.S.; Mohamed, J.S.; Hollander, J.M.; Alway, S.E. The role of SIRT1 in skeletal muscle function and repair of older mice. J. Cachexia Sarcopenia Muscle 2019, 10, 929–949. [Google Scholar] [CrossRef] [Green Version]
  91. Suetta, C.; Frandsen, U.; Mackey, A.L.; Jensen, L.; Hvid, L.G.; Bayer, M.; Petersson, S.J.; Schrøder, H.D.; Andersen, J.L.; Aagaard, P. Ageing is associated with diminished muscle re-growth and myogenic precursor cell expansion early after immobility-induced atrophy in human skeletal muscle. J. Physiol. 2013, 591, 3789–3804. [Google Scholar] [CrossRef]
  92. Ahmadi, M.; Karlsen, A.; Mehling, J.; Soendenbroe, C.; Mackey, A.L.; Hyldahl, R.D. Aging is associated with an altered macrophage response during human skeletal muscle regeneration. Exp. Gerontol. 2022, 169, 111974. [Google Scholar] [CrossRef]
  93. Phillips, S.M.; McGlory, C. CrossTalk proposal: The dominant mechanism causing disuse muscle atrophy is decreased protein synthesis. J. Physiol. 2014, 592, 5341. [Google Scholar] [CrossRef]
  94. Alway, S.E.; Bennett, B.T.; Wilson, J.C.; Sperringer, J.; Mohamed, J.S.; Edens, N.K.; Pereira, S.L. Green tea extract attenuates muscle loss and improves muscle function during disuse, but fails to improve muscle recovery following unloading in aged rats. J. Appl. Physiol. 2015, 118, 319–330. [Google Scholar] [CrossRef] [Green Version]
  95. Siu, P.M.; Pistilli, E.E.; Alway, S.E. Age-dependent increase in oxidative stress in gastrocnemius muscle with unloading. J. Appl. Physiol. 2008, 105, 1695–1705. [Google Scholar] [CrossRef] [Green Version]
  96. Siu, P.M.; Pistilli, E.E.; Murlasits, Z.; Alway, S.E. Hindlimb unloading increases muscle content of cytosolic but not nuclear Id2 and p53 proteins in young adult and aged rats. J. Appl. Physiol. 2006, 100, 907–916. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Siu, P.M.; Pistilli, E.E.; Butler, D.C.; Alway, S.E. Aging influences cellular and molecular responses of apoptosis to skeletal muscle unloading. Am. J. Physiol.-Cell Physiol. 2005, 288, C338–C349. [Google Scholar] [CrossRef] [PubMed]
  98. Siu, P.M.; Pistilli, E.E.; Alway, S.E. Apoptotic responses to hindlimb suspension in gastrocnemius muscles from young adult and aged rats. Am. J. Physiol.-Regul. Integr. Comp. Physiol. 2005, 289, R1015–R1026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Oliveira, J.R.S.; Mohamed, J.S.; Myers, M.J.; Brooks, M.J.; Alway, S.E. Effects of hindlimb suspension and reloading on gastrocnemius and soleus muscle mass and function in geriatric mice. Exp. Gerontol. 2019, 115, 19–31. [Google Scholar] [CrossRef]
  100. Bennett, B.T.; Mohamed, J.S.; Alway, S.E. Effects of resveratrol on the recovery of muscle mass following disuse in the plantaris muscle of aged rats. PLoS ONE 2013, 8, e83518. [Google Scholar] [CrossRef] [Green Version]
  101. Hao, Y.; Jackson, J.R.; Wang, Y.; Edens, N.; Pereira, S.L.; Alway, S.E. β-Hydroxy-β-methylbutyrate reduces myonuclear apoptosis during recovery from hind limb suspension-induced muscle fiber atrophy in aged rats. Am. J. Physiol.-Regul. Integr. Comp. Physiol. 2011, 301, R701–R715. [Google Scholar] [CrossRef] [Green Version]
  102. Zhang, X.; Trevino, M.B.; Wang, M.; Gardell, S.J.; Ayala, J.E.; Han, X.; Kelly, D.P.; Goodpaster, B.H.; Vega, R.B.; Coen, P.M. Impaired mitochondrial energetics characterize poor early recovery of muscle mass following hind limb unloading in old mice. J. Gerontol. Ser. A 2018, 73, 1313–1322. [Google Scholar] [CrossRef] [Green Version]
  103. Kirby, T.J.; Lee, J.D.; England, J.H.; Chaillou, T.; Esser, K.A.; McCarthy, J.J. Blunted hypertrophic response in aged skeletal muscle is associated with decreased ribosome biogenesis. J. Appl. Physiol. 2015, 119, 321–327. [Google Scholar] [CrossRef]
  104. Trevino, M.B.; Zhang, X.; Standley, R.A.; Wang, M.; Han, X.; Reis, F.C.; Periasamy, M.; Yu, G.; Kelly, D.P.; Goodpaster, B.H. Loss of mitochondrial energetics is associated with poor recovery of muscle function but not mass following disuse atrophy. Am. J. Physiol.-Endocrinol. Metab. 2019, 317, E899–E910. [Google Scholar] [CrossRef] [Green Version]
  105. Baehr, L.M.; West, D.W.; Marcotte, G.; Marshall, A.G.; De Sousa, L.G.; Baar, K.; Bodine, S.C. Age-related deficits in skeletal muscle recovery following disuse are associated with neuromuscular junction instability and ER stress, not impaired protein synthesis. Aging (Albany NY) 2016, 8, 127. [Google Scholar] [CrossRef] [Green Version]
  106. Baehr, L.M.; West, D.W.; Marshall, A.G.; Marcotte, G.R.; Baar, K.; Bodine, S.C. Muscle-specific and age-related changes in protein synthesis and protein degradation in response to hindlimb unloading in rats. J. Appl. Physiol. 2017, 122, 1336–1350. [Google Scholar] [CrossRef] [Green Version]
  107. Waterlow, J. Protein turnover with special reference to man. Q. J. Exp. Physiol. Transl. Integr. 1984, 69, 409–438. [Google Scholar] [CrossRef]
  108. Rolfe, D.; Brown, G.C. Cellular energy utilization and molecular origin of standard metabolic rate in mammals. Physiol. Rev. 1997, 77, 731–758. [Google Scholar] [CrossRef] [Green Version]
  109. Liu, J.; Peng, Y.; Feng, Z.; Shi, W.; Qu, L.; Li, Y.; Liu, J.; Long, J. Reloading functionally ameliorates disuse-induced muscle atrophy by reversing mitochondrial dysfunction, and similar benefits are gained by administering a combination of mitochondrial nutrients. Free Radic. Biol. Med. 2014, 69, 116–128. [Google Scholar] [CrossRef]
  110. Hood, D.A.; Memme, J.M.; Oliveira, A.N.; Triolo, M. Maintenance of Skeletal Muscle Mitochondria in Health, Exercise, and Aging. Annu. Rev. Physiol. 2019, 81, 19–41. [Google Scholar] [CrossRef]
  111. Brooks, M.J.; Hajira, A.; Mohamed, J.S.; Alway, S.E. Voluntary wheel running increases satellite cell abundance and improves recovery from disuse in gastrocnemius muscles from mice. J. Appl. Physiol. 2018, 124, 1616–1628. [Google Scholar] [CrossRef]
  112. Hwee, D.T.; Bodine, S.C. Age-related deficit in load-induced skeletal muscle growth. J. Gerontol. Ser. A Biomed. Sci. Med. Sci. 2009, 64, 618–628. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Suetta, C.; Hvid, L.G.; Justesen, L.; Christensen, U.; Neergaard, K.; Simonsen, L.; Ortenblad, N.; Magnusson, S.P.; Kjaer, M.; Aagaard, P. Effects of aging on human skeletal muscle after immobilization and retraining. J. Appl. Physiol. 2009, 107, 1172–1180. [Google Scholar] [CrossRef] [PubMed]
  114. Tanner, R.E.; Brunker, L.B.; Agergaard, J.; Barrows, K.M.; Briggs, R.A.; Kwon, O.S.; Young, L.M.; Hopkins, P.N.; Volpi, E.; Marcus, R.L. Age-related differences in lean mass, protein synthesis and skeletal muscle markers of proteolysis after bed rest and exercise rehabilitation. J. Physiol. 2015, 593, 4259–4273. [Google Scholar] [CrossRef] [Green Version]
  115. Sandri, M.; Barberi, L.; Bijlsma, A.; Blaauw, B.; Dyar, K.; Milan, G.; Mammucari, C.; Meskers, C.; Pallafacchina, G.; Paoli, A. Signalling pathways regulating muscle mass in ageing skeletal muscle. The role of the IGF1-Akt-mTOR-FoxO pathway. Biogerontology 2013, 14, 303–323. [Google Scholar] [CrossRef] [PubMed]
  116. White, Z.; White, R.B.; McMahon, C.; Grounds, M.D.; Shavlakadze, T. High mTORC1 signaling is maintained, while protein degradation pathways are perturbed in old murine skeletal muscles in the fasted state. Int. J. Biochem. Cell Biol. 2016, 78, 10–21. [Google Scholar] [CrossRef]
  117. Joseph, G.A.; Wang, S.X.; Jacobs, C.E.; Zhou, W.; Kimble, G.C.; Tse, H.W.; Eash, J.K.; Shavlakadze, T.; Glass, D.J. Partial inhibition of mTORC1 in aged rats counteracts the decline in muscle mass and reverses molecular signaling associated with sarcopenia. Mol. Cell. Biol. 2019, 39, e00141-19. [Google Scholar] [CrossRef] [Green Version]
  118. Markofski, M.M.; Dickinson, J.M.; Drummond, M.J.; Fry, C.S.; Fujita, S.; Gundermann, D.M.; Glynn, E.L.; Jennings, K.; Paddon-Jones, D.; Reidy, P.T. Effect of age on basal muscle protein synthesis and mTORC1 signaling in a large cohort of young and older men and women. Exp. Gerontol. 2015, 65, 1–7. [Google Scholar] [CrossRef] [Green Version]
  119. Baar, E.L.; Carbajal, K.A.; Ong, I.M.; Lamming, D.W. Sex-and tissue-specific changes in mTOR signaling with age in C57 BL/6J mice. Aging Cell 2016, 15, 155–166. [Google Scholar] [CrossRef]
  120. Blaauw, B.; Canato, M.; Agatea, L.; Toniolo, L.; Mammucari, C.; Masiero, E.; Abraham, R.; Sandri, M.; Schiaffino, S.; Reggiani, C. Inducible activation of Akt increases skeletal muscle mass and force without satellite cell activation. FASEB J. 2009, 23, 3896–3905. [Google Scholar] [CrossRef]
  121. Bentzinger, C.F.; Lin, S.; Romanino, K.; Castets, P.; Guridi, M.; Summermatter, S.; Handschin, C.; Tintignac, L.A.; Hall, M.N.; Rüegg, M.A. Differential response of skeletal muscles to mTORC1 signaling during atrophy and hypertrophy. Skelet. Muscle 2013, 3, 6. [Google Scholar] [CrossRef] [Green Version]
  122. Tang, H.; Inoki, K.; Brooks, S.V.; Okazawa, H.; Lee, M.; Wang, J.; Kim, M.; Kennedy, C.L.; Macpherson, P.C.; Ji, X. mTORC1 underlies age-related muscle fiber damage and loss by inducing oxidative stress and catabolism. Aging Cell 2019, 18, e12943. [Google Scholar] [CrossRef] [Green Version]
  123. Guridi, M.; Tintignac, L.A.; Lin, S.; Kupr, B.; Castets, P.; Rüegg, M.A. Activation of mTORC1 in skeletal muscle regulates whole-body metabolism through FGF21. Sci. Signal. 2015, 8, ra113. [Google Scholar] [CrossRef]
  124. Bentzinger, C.F.; Romanino, K.; Cloëtta, D.; Lin, S.; Mascarenhas, J.B.; Oliveri, F.; Xia, J.; Casanova, E.; Costa, C.F.; Brink, M. Skeletal muscle-specific ablation of raptor, but not of rictor, causes metabolic changes and results in muscle dystrophy. Cell Metab. 2008, 8, 411–424. [Google Scholar] [CrossRef] [Green Version]
  125. Ham, D.J.; Börsch, A.; Chojnowska, K.; Lin, S.; Leuchtmann, A.B.; Ham, A.S.; Thürkauf, M.; Delezie, J.; Furrer, R.; Burri, D. Distinct and additive effects of calorie restriction and rapamycin in aging skeletal muscle. Nat. Commun. 2022, 13, 2025. [Google Scholar] [CrossRef]
  126. Alway, S.E.; Mohamed, J.S.; Myers, M.J. Mitochondria initiate and regulate sarcopenia. Exerc. Sport. Sci. Rev. 2017, 45, 58–69. [Google Scholar] [CrossRef] [Green Version]
  127. Triolo, M.; Oliveira, A.N.; Kumari, R.; Hood, D.A. The influence of age, sex, and exercise on autophagy, mitophagy, and lysosome biogenesis in skeletal muscle. Skelet. Muscle 2022, 12, 13. [Google Scholar] [CrossRef]
  128. Kang, C.; Yeo, D.; Ji, L. Muscle immobilization activates mitophagy and disrupts mitochondrial dynamics in mice. Acta Physiol. 2016, 218, 188–197. [Google Scholar] [CrossRef]
  129. Bard, J.A.; Goodall, E.A.; Greene, E.R.; Jonsson, E.; Dong, K.C.; Martin, A. Structure and function of the 26S proteasome. Annu. Rev. Biochem. 2018, 87, 697. [Google Scholar] [CrossRef]
  130. Ciechanover, A.; Stanhill, A. The complexity of recognition of ubiquitinated substrates by the 26S proteasome. Biochim. Et Biophys. Acta (BBA)-Mol. Cell Res. 2014, 1843, 86–96. [Google Scholar] [CrossRef] [Green Version]
  131. Korovila, I.; Hugo, M.; Castro, J.P.; Weber, D.; Höhn, A.; Grune, T.; Jung, T. Proteostasis, oxidative stress and aging. Redox Biol. 2017, 13, 550–567. [Google Scholar] [CrossRef]
  132. Hughes, D.C.; Baehr, L.M.; Waddell, D.S.; Sharples, A.P.; Bodine, S.C. Ubiquitin Ligases in Longevity and Aging Skeletal Muscle. Int. J. Mol. Sci. 2022, 23, 7602. [Google Scholar] [CrossRef] [PubMed]
  133. Fernando, R.; Drescher, C.; Nowotny, K.; Grune, T.; Castro, J.P. Impaired proteostasis during skeletal muscle aging. Free Radic. Biol. Med. 2019, 132, 58–66. [Google Scholar] [CrossRef] [PubMed]
  134. O’Neill, B.T.; Bhardwaj, G.; Penniman, C.M.; Krumpoch, M.T.; Suarez Beltran, P.A.; Klaus, K.; Poro, K.; Li, M.; Pan, H.; Dreyfuss, J.M. FoxO transcription factors are critical regulators of diabetes-related muscle atrophy. Diabetes 2019, 68, 556–570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Reich, K.A.; Chen, Y.-W.; Thompson, P.D.; Hoffman, E.P.; Clarkson, P.M. Forty-eight hours of unloading and 24 h of reloading lead to changes in global gene expression patterns related to ubiquitination and oxidative stress in humans. J. Appl. Physiol. 2010, 109, 1404–1415. [Google Scholar] [CrossRef] [PubMed]
  136. Caron, A.Z.; Haroun, S.; Leblanc, É.; Trensz, F.; Guindi, C.; Amrani, A.; Grenier, G. The proteasome inhibitor MG132 reduces immobilization-induced skeletal muscle atrophy in mice. BMC Musculoskelet. Disord. 2011, 12, 185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Bodine, S.C.; Latres, E.; Baumhueter, S.; Lai, V.K.-M.; Nunez, L.; Clarke, B.A.; Poueymirou, W.T.; Panaro, F.J.; Na, E.; Dharmarajan, K. Identification of ubiquitin ligases required for skeletal muscle atrophy. Science 2001, 294, 1704–1708. [Google Scholar] [CrossRef]
  138. Sandri, M.; Lin, J.; Handschin, C.; Yang, W.; Arany, Z.P.; Lecker, S.H.; Goldberg, A.L.; Spiegelman, B.M. PGC-1α protects skeletal muscle from atrophy by suppressing FoxO3 action and atrophy-specific gene transcription. Proc. Natl. Acad. Sci. USA 2006, 103, 16260–16265. [Google Scholar] [CrossRef] [Green Version]
  139. Brocca, L.; Toniolo, L.; Reggiani, C.; Bottinelli, R.; Sandri, M.; Pellegrino, M.A. FoxO-dependent atrogenes vary among catabolic conditions and play a key role in muscle atrophy induced by hindlimb suspension. J. Physiol. 2017, 595, 1143–1158. [Google Scholar] [CrossRef] [Green Version]
  140. Jagoe, R.T.; Lecker, S.H.; Gomes, M.; Goldberg, A.L. Patterns of gene expression in atrophying skeletal muscles: Response to food deprivation. FASEB J. 2002, 16, 1697–1712. [Google Scholar] [CrossRef] [Green Version]
  141. Lagirand-Cantaloube, J.; Cornille, K.; Csibi, A.; Batonnet-Pichon, S.; Leibovitch, M.P.; Leibovitch, S.A. Inhibition of atrogin-1/MAFbx mediated MyoD proteolysis prevents skeletal muscle atrophy in vivo. PLoS ONE 2009, 4, e4973. [Google Scholar] [CrossRef] [Green Version]
  142. Tintignac, L.A.; Lagirand, J.; Batonnet, S.; Sirri, V.; Leibovitch, M.P.; Leibovitch, S.A. Degradation of MyoD mediated by the SCF (MAFbx) ubiquitin ligase. J. Biol. Chem. 2005, 280, 2847–2856. [Google Scholar] [CrossRef] [Green Version]
  143. Jogo, M.; Shiraishi, S.; Tamura, T.-A. Identification of MAFbx as a myogenin-engaged F-box protein in SCF ubiquitin ligase. FEBS Lett. 2009, 583, 2715–2719. [Google Scholar] [CrossRef]
  144. Csibi, A.; Leibovitch, M.P.; Cornille, K.; Tintignac, L.A.; Leibovitch, S.A. MAFbx/Atrogin-1 controls the activity of the initiation factor eIF3-f in skeletal muscle atrophy by targeting multiple C-terminal lysines. J. Biol. Chem. 2009, 284, 4413–4421. [Google Scholar] [CrossRef] [Green Version]
  145. Lokireddy, S.; Wijesoma, I.W.; Sze, S.K.; McFarlane, C.; Kambadur, R.; Sharma, M. Identification of atrogin-1-targeted proteins during the myostatin-induced skeletal muscle wasting. Am. J. Physiol.-Cell Physiol. 2012, 303, C512–C529. [Google Scholar] [CrossRef] [Green Version]
  146. Clarke, B.A.; Drujan, D.; Willis, M.S.; Murphy, L.O.; Corpina, R.A.; Burova, E.; Rakhilin, S.V.; Stitt, T.N.; Patterson, C.; Latres, E. The E3 Ligase MuRF1 degrades myosin heavy chain protein in dexamethasone-treated skeletal muscle. Cell Metab. 2007, 6, 376–385. [Google Scholar] [CrossRef] [Green Version]
  147. Baehr, L.M.; Hughes, D.C.; Lynch, S.A.; Van Haver, D.; Maia, T.M.; Marshall, A.G.; Radoshevich, L.; Impens, F.; Waddell, D.S.; Bodine, S.C. Identification of the MuRF1 skeletal muscle ubiquitylome through quantitative proteomics. Function 2021, 2, zqab029. [Google Scholar] [CrossRef]
  148. Rudolf, R.; Bogomolovas, J.; Strack, S.; Choi, K.-R.; Khan, M.M.; Wagner, A.; Brohm, K.; Hanashima, A.; Gasch, A.; Labeit, D. Regulation of nicotinic acetylcholine receptor turnover by MuRF1 connects muscle activity to endo/lysosomal and atrophy pathways. Age 2013, 35, 1663–1674. [Google Scholar] [CrossRef] [Green Version]
  149. 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]
  150. Stitt, T.N.; Drujan, D.; Clarke, B.A.; Panaro, F.; Timofeyva, Y.; Kline, W.O.; Gonzalez, M.; Yancopoulos, G.D.; Glass, D.J. The IGF-1/PI3K/Akt pathway prevents expression of muscle atrophy-induced ubiquitin ligases by inhibiting FOXO transcription factors. Mol. Cell 2004, 14, 395–403. [Google Scholar] [CrossRef]
  151. 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]
  152. Sanchez, A.M.; Csibi, A.; Raibon, A.; Cornille, K.; Gay, S.; Bernardi, H.; Candau, R. AMPK promotes skeletal muscle autophagy through activation of forkhead FoxO3a and interaction with Ulk1. J. Cell. Biochem. 2012, 113, 695–710. [Google Scholar] [CrossRef] [PubMed]
  153. Greer, E.L.; Oskoui, P.R.; Banko, M.R.; Maniar, J.M.; Gygi, M.P.; Gygi, S.P.; Brunet, A. The energy sensor AMP-activated protein kinase directly regulates the mammalian FOXO3 transcription factor. J. Biol. Chem. 2007, 282, 30107–30119. [Google Scholar] [CrossRef] [PubMed]
  154. Martinez-Vicente, M.; Sovak, G.; Cuervo, A.M. Protein degradation and aging. Exp. Gerontol. 2005, 40, 622–633. [Google Scholar] [CrossRef] [PubMed]
  155. Kitajima, Y.; Tashiro, Y.; Suzuki, N.; Warita, H.; Kato, M.; Tateyama, M.; Ando, R.; Izumi, R.; Yamazaki, M.; Abe, M. Proteasome dysfunction induces muscle growth defects and protein aggregation. J. Cell Sci. 2014, 127, 5204–5217. [Google Scholar] [PubMed] [Green Version]
  156. Fielitz, J.; Kim, M.-S.; Shelton, J.M.; Latif, S.; Spencer, J.A.; Glass, D.J.; Richardson, J.A.; Bassel-Duby, R.; Olson, E.N. Myosin accumulation and striated muscle myopathy result from the loss of muscle RING finger 1 and 3. J. Clin. Investig. 2007, 117, 2486–2495. [Google Scholar] [CrossRef]
  157. Gomes, A.V.; Waddell, D.S.; Siu, R.; Stein, M.; Dewey, S.; Furlow, J.D.; Bodine, S.C. Upregulation of proteasome activity in muscle RING finger 1-null mice following denervation. FASEB J. 2012, 26, 2986–2999. [Google Scholar] [CrossRef] [Green Version]
  158. Hwee, D.T.; Baehr, L.M.; Philp, A.; Baar, K.; Bodine, S.C. Maintenance of muscle mass and load-induced growth in Muscle RING Finger 1 null mice with age. Aging Cell 2014, 13, 92–101. [Google Scholar] [CrossRef]
  159. Altun, M.; Besche, H.C.; Overkleeft, H.S.; Piccirillo, R.; Edelmann, M.J.; Kessler, B.M.; Goldberg, A.L.; Ulfhake, B. Muscle wasting in aged, sarcopenic rats is associated with enhanced activity of the ubiquitin proteasome pathway. J. Biol. Chem. 2010, 285, 39597–39608. [Google Scholar] [CrossRef] [Green Version]
  160. Raue, U.; Slivka, D.; Jemiolo, B.; Hollon, C.; Trappe, S. Proteolytic gene expression differs at rest and after resistance exercise between young and old women. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 2007, 62, 1407–1412. [Google Scholar] [CrossRef] [Green Version]
  161. Clavel, S.; Coldefy, A.-S.; Kurkdjian, E.; Salles, J.; Margaritis, I.; Derijard, B. Atrophy-related ubiquitin ligases, atrogin-1 and MuRF1 are up-regulated in aged rat Tibialis Anterior muscle. Mech. Ageing Dev. 2006, 127, 794–801. [Google Scholar] [CrossRef]
  162. Strucksberg, K.-H.; Tangavelou, K.; Schröder, R.; Clemen, C.S. Proteasomal activity in skeletal muscle: A matter of assay design, muscle type, and age. Anal. Biochem. 2010, 399, 225–229. [Google Scholar] [CrossRef]
  163. Edström, E.; Altun, M.; Hägglund, M.; Ulfhake, B. Atrogin-1/MAFbx and MuRF1 are downregulated in aging-related loss of skeletal muscle. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 2006, 61, 663–674. [Google Scholar] [CrossRef]
  164. Kaiser, M.S.; Milan, G.; Ham, D.J.; Lin, S.; Oliveri, F.; Chojnowska, K.; Tintignac, L.A.; Mittal, N.; Zimmerli, C.E.; Glass, D.J.; et al. Dual roles of mTORC1-dependent activation of the ubiquitin-proteasome system in muscle proteostasis. Commun. Biol. 2022, 5, 1141. [Google Scholar] [CrossRef]
  165. Kido, K.; Sase, K.; Yokokawa, T.; Fujita, S. Enhanced skeletal muscle insulin sensitivity after acute resistance-type exercise is upregulated by rapamycin-sensitive mTOR complex 1 inhibition. Sci. Rep. 2020, 10, 8509. [Google Scholar] [CrossRef]
  166. Ogata, T.; Oishi, Y.; Higuchi, M.; Muraoka, I. Fasting-related autophagic response in slow-and fast-twitch skeletal muscle. Biochem. Biophys. Res. Commun. 2010, 394, 136–140. [Google Scholar] [CrossRef]
  167. 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] [Green Version]
  168. Vainshtein, A.; Hood, D.A. The regulation of autophagy during exercise in skeletal muscle. J. Appl. Physiol. 2016, 120, 664–673. [Google Scholar] [CrossRef] [Green Version]
  169. McClung, J.M.; Judge, A.R.; Powers, S.K.; Yan, Z. p38 MAPK links oxidative stress to autophagy-related gene expression in cachectic muscle wasting. Am. J. Physiol.-Cell Physiol. 2010, 298, C542–C549. [Google Scholar] [CrossRef] [Green Version]
  170. Madaro, L.; Marrocco, V.; Carnio, S.; Sandri, M.; Bouché, M. Intracellular signaling in ER stress-induced autophagy in skeletal muscle cells. FASEB J. 2013, 27, 1990–2000. [Google Scholar] [CrossRef] [Green Version]
  171. Penna, F.; Costamagna, D.; Pin, F.; Camperi, A.; Fanzani, A.; Chiarpotto, E.M.; Cavallini, G.; Bonelli, G.; Baccino, F.M.; Costelli, P. Autophagic degradation contributes to muscle wasting in cancer cachexia. Am. J. Pathol. 2013, 182, 1367–1378. [Google Scholar] [CrossRef]
  172. Hosokawa, N.; Hara, T.; Kaizuka, T.; Kishi, C.; Takamura, A.; Miura, Y.; Iemura, S.-I.; Natsume, T.; Takehana, K.; Yamada, N. Nutrient-dependent mTORC1 association with the ULK1–Atg13–FIP200 complex required for autophagy. Mol. Biol. Cell 2009, 20, 1981–1991. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Hara, T.; Takamura, A.; Kishi, C.; Iemura, S.-I.; Natsume, T.; Guan, J.-L.; Mizushima, N. FIP200, a ULK-interacting protein, is required for autophagosome formation in mammalian cells. J. Cell Biol. 2008, 181, 497–510. [Google Scholar] [CrossRef] [PubMed]
  174. Mercer, C.A.; Kaliappan, A.; Dennis, P.B. A novel, human Atg13 binding protein, Atg101, interacts with ULK1 and is essential for macroautophagy. Autophagy 2009, 5, 649–662. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Jung, C.H.; Jun, C.B.; Ro, S.-H.; Kim, Y.-M.; Otto, N.M.; Cao, J.; Kundu, M.; Kim, D.-H. ULK-Atg13-FIP200 complexes mediate mTOR signaling to the autophagy machinery. Mol. Biol. Cell 2009, 20, 1992–2003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Kim, J.; Kundu, M.; Viollet, B.; Guan, K.-L. AMPK and mTOR regulate autophagy through direct phosphorylation of Ulk1. Nat. Cell Biol. 2011, 13, 132–141. [Google Scholar] [CrossRef] [Green Version]
  177. Nazio, F.; Strappazzon, F.; Antonioli, M.; Bielli, P.; Cianfanelli, V.; Bordi, M.; Gretzmeier, C.; Dengjel, J.; Piacentini, M.; Fimia, G.M. mTOR inhibits autophagy by controlling ULK1 ubiquitylation, self-association and function through AMBRA1 and TRAF6. Nat. Cell Biol. 2013, 15, 406–416. [Google Scholar] [CrossRef]
  178. Di Bartolomeo, S.; Corazzari, M.; Nazio, F.; Oliverio, S.; Lisi, G.; Antonioli, M.; Pagliarini, V.; Matteoni, S.; Fuoco, C.; Giunta, L. The dynamic interaction of AMBRA1 with the dynein motor complex regulates mammalian autophagy. J. Cell Biol. 2010, 191, 155–168. [Google Scholar] [CrossRef] [Green Version]
  179. Russell, R.C.; Tian, Y.; Yuan, H.; Park, H.W.; Chang, Y.-Y.; Kim, J.; Kim, H.; Neufeld, T.P.; Dillin, A.; Guan, K.-L. ULK1 induces autophagy by phosphorylating Beclin-1 and activating VPS34 lipid kinase. Nat. Cell Biol. 2013, 15, 741–750. [Google Scholar] [CrossRef] [Green Version]
  180. Proikas-Cezanne, T.; Takacs, Z.; Dönnes, P.; Kohlbacher, O. WIPI proteins: Essential PtdIns3 P effectors at the nascent autophagosome. J. Cell Sci. 2015, 128, 207–217. [Google Scholar] [CrossRef] [Green Version]
  181. Nakatogawa, H.; Ichimura, Y.; Ohsumi, Y. Atg8, a ubiquitin-like protein required for autophagosome formation, mediates membrane tethering and hemifusion. Cell 2007, 130, 165–178. [Google Scholar] [CrossRef] [Green Version]
  182. Sou, Y.-S.; Waguri, S.; Iwata, J.-I.; Ueno, T.; Fujimura, T.; Hara, T.; Sawada, N.; Yamada, A.; Mizushima, N.; Uchiyama, Y. The Atg8 conjugation system is indispensable for proper development of autophagic isolation membranes in mice. Mol. Biol. Cell 2008, 19, 4762–4775. [Google Scholar] [CrossRef] [Green Version]
  183. Weidberg, H.; Shvets, E.; Shpilka, T.; Shimron, F.; Shinder, V.; Elazar, Z. LC3 and GATE-16/GABARAP subfamilies are both essential yet act differently in autophagosome biogenesis. EMBO J. 2010, 29, 1792–1802. [Google Scholar] [CrossRef] [Green Version]
  184. Fallaize, D.; Chin, L.-S.; Li, L. Differential submitochondrial localization of PINK1 as a molecular switch for mediating distinct mitochondrial signaling pathways. Cell. Signal. 2015, 27, 2543–2554. [Google Scholar] [CrossRef] [Green Version]
  185. 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]
  186. 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]
  187. Sarraf, S.A.; Raman, M.; Guarani-Pereira, V.; Sowa, M.E.; Huttlin, E.L.; Gygi, S.P.; Harper, J.W. Landscape of the PARKIN-dependent ubiquitylome in response to mitochondrial depolarization. Nature 2013, 496, 372–376. [Google Scholar] [CrossRef] [Green Version]
  188. Novak, I.; Kirkin, V.; McEwan, D.G.; Zhang, J.; Wild, P.; Rozenknop, A.; Rogov, V.; Löhr, F.; Popovic, D.; Occhipinti, A. Nix is a selective autophagy receptor for mitochondrial clearance. EMBO Rep. 2010, 11, 45–51. [Google Scholar] [CrossRef] [Green Version]
  189. Lazarou, M.; Sliter, D.A.; Kane, L.A.; Sarraf, S.A.; Wang, C.; Burman, J.L.; Sideris, D.P.; Fogel, A.I.; Youle, R.J. The ubiquitin kinase PINK1 recruits autophagy receptors to induce mitophagy. Nature 2015, 524, 309–314. [Google Scholar] [CrossRef] [Green Version]
  190. Wohlgemuth, S.E.; Seo, A.Y.; Marzetti, E.; Lees, H.A.; Leeuwenburgh, C. Skeletal muscle autophagy and apoptosis during aging: Effects of calorie restriction and life-long exercise. Exp. Gerontol. 2010, 45, 138–148. [Google Scholar] [CrossRef] [Green Version]
  191. 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] [Green Version]
  192. 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]
  193. 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] [PubMed]
  194. Fry, C.S.; Drummond, M.J.; Glynn, E.L.; Dickinson, J.M.; Gundermann, D.M.; Timmerman, K.L.; Walker, D.K.; Volpi, E.; Rasmussen, B.B. Skeletal muscle autophagy and protein breakdown following resistance exercise are similar in younger and older adults. J. Gerontol. Ser. A Biomed. Sci. Med. Sci. 2013, 68, 599–607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Carnio, S.; LoVerso, F.; Baraibar, M.A.; Longa, E.; Khan, M.M.; Maffei, M.; Reischl, M.; Canepari, M.; Loefler, S.; Kern, H. Autophagy impairment in muscle induces neuromuscular junction degeneration and precocious aging. Cell Rep. 2014, 8, 1509–1521. [Google Scholar] [CrossRef] [PubMed]
  196. 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] [Green Version]
  197. 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 (Albany NY) 2020, 12, 5977. [Google Scholar] [CrossRef]
  198. Mizushima, N.; Yamamoto, A.; Matsui, M.; Yoshimori, T.; Ohsumi, Y. In vivo analysis of autophagy in response to nutrient starvation using transgenic mice expressing a fluorescent autophagosome marker. Mol. Biol. Cell 2004, 15, 1101–1111. [Google Scholar] [CrossRef] [Green Version]
  199. White, Z.; Terrill, J.; White, R.B.; McMahon, C.; Sheard, P.; Grounds, M.D.; Shavlakadze, T. Voluntary resistance wheel exercise from mid-life prevents sarcopenia and increases markers of mitochondrial function and autophagy in muscles of old male and female C57BL/6J mice. Skelet. Muscle 2016, 6, 45. [Google Scholar] [CrossRef] [Green Version]
  200. Lira, V.A.; Okutsu, M.; Zhang, M.; Greene, N.P.; Laker, R.C.; Breen, D.S.; Hoehn, K.L.; Yan, Z. Autophagy is required for exercise training-induced skeletal muscle adaptation and improvement of physical performance. FASEB J. 2013, 27, 4184–4193. [Google Scholar] [CrossRef] [Green Version]
  201. 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]
  202. Bujak, A.L.; Crane, J.D.; Lally, J.S.; Ford, R.J.; Kang, S.J.; Rebalka, I.A.; Green, A.E.; Kemp, B.E.; Hawke, T.J.; Schertzer, J.D. AMPK activation of muscle autophagy prevents fasting-induced hypoglycemia and myopathy during aging. Cell Metab. 2015, 21, 883–890. [Google Scholar] [CrossRef] [Green Version]
  203. Huang, Y.; Zhu, X.; Chen, K.; Lang, H.; Zhang, Y.; Hou, P.; Ran, L.; Zhou, M.; Zheng, J.; Yi, L. Resveratrol prevents sarcopenic obesity by reversing mitochondrial dysfunction and oxidative stress via the PKA/LKB1/AMPK pathway. Aging 2019, 11, 2217. [Google Scholar] [CrossRef]
  204. 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] [Green Version]
  205. Penniman, C.M.; Bhardwaj, G.; Nowers, C.J.; Brown, C.U.; Junck, T.L.; Boyer, C.K.; Jena, J.; Fuqua, J.D.; Lira, V.A.; O’Neill, B.T. Loss of FoxOs in muscle increases strength and mitochondrial function during aging. J. Cachexia Sarcopenia Muscle 2022. [Google Scholar] [CrossRef]
  206. Castets, P.; Lin, S.; Rion, N.; Di Fulvio, S.; Romanino, K.; Guridi, M.; Frank, S.; Tintignac, L.A.; Sinnreich, M.; Rüegg, M.A. Sustained activation of mTORC1 in skeletal muscle inhibits constitutive and starvation-induced autophagy and causes a severe, late-onset myopathy. Cell Metab. 2013, 17, 731–744. [Google Scholar] [CrossRef] [Green Version]
  207. Crombie, E.M.; Kim, S.; Adamson, S.; Dong, H.; Lu, T.C.; Wu, Y.; Wu, Y.; Levy, Y.; Stimple, N.; Lam, W.M.R. Activation of eIF4E-binding-protein-1 rescues mTORC1-induced sarcopenia by expanding lysosomal degradation capacity. J. Cachexia Sarcopenia Muscle 2022. [Google Scholar] [CrossRef]
  208. Picard, M.; Ritchie, D.; Wright, K.J.; Romestaing, C.; Thomas, M.M.; Rowan, S.L.; Taivassalo, T.; Hepple, R.T. Mitochondrial functional impairment with aging is exaggerated in isolated mitochondria compared to permeabilized myofibers. Aging Cell 2010, 9, 1032–1046. [Google Scholar] [CrossRef] [Green Version]
  209. Picard, M.; Taivassalo, T.; Ritchie, D.; Wright, K.J.; Thomas, M.M.; Romestaing, C.; Hepple, R.T. Mitochondrial structure and function are disrupted by standard isolation methods. PLoS ONE 2011, 6, e18317. [Google Scholar] [CrossRef]
  210. Gouspillou, G.; Sgarioto, N.; Kapchinsky, S.; Purves-Smith, F.; Norris, B.; Pion, C.H.; Barbat-Artigas, S.; Lemieux, F.; Taivassalo, T.; Morais, J.A. 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]
  211. Terman, A.; Kurz, T.; Navratil, M.; Arriaga, E.A.; Brunk, U.T. Mitochondrial turnover and aging of long-lived postmitotic cells: The mitochondrial–lysosomal axis theory of aging. Antioxid. Redox Signal. 2010, 12, 503–535. [Google Scholar] [CrossRef] [Green Version]
  212. Holloway, G.P.; Holwerda, A.M.; Miotto, P.M.; Dirks, M.L.; Verdijk, L.B.; van Loon, L.J. Age-associated impairments in mitochondrial ADP sensitivity contribute to redox stress in senescent human skeletal muscle. Cell Rep. 2018, 22, 2837–2848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Merz, K.E.; Hwang, J.; Zhou, C.; Veluthakal, R.; McCown, E.M.; Hamilton, A.; Oh, E.; Dai, W.; Fueger, P.T.; Jiang, L. Enrichment of the exocytosis protein STX4 in skeletal muscle remediates peripheral insulin resistance and alters mitochondrial dynamics via Drp1. Nat. Commun. 2022, 13, 424. [Google Scholar] [CrossRef] [PubMed]
  214. Miotto, P.M.; LeBlanc, P.J.; Holloway, G.P. High-fat diet causes mitochondrial dysfunction as a result of impaired ADP sensitivity. Diabetes 2018, 67, 2199–2205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Dantas, W.S.; Zunica, E.R.; Heintz, E.C.; Vandanmagsar, B.; Floyd, Z.E.; Yu, Y.; Fujioka, H.; Hoppel, C.L.; Belmont, K.P.; Axelrod, C.L. Mitochondrial uncoupling attenuates sarcopenic obesity by enhancing skeletal muscle mitophagy and quality control. J. Cachexia Sarcopenia Muscle 2022, 13, 1821–1836. [Google Scholar] [CrossRef] [PubMed]
  216. Andersson, D.C.; Betzenhauser, M.J.; Reiken, S.; Meli, A.C.; Umanskaya, A.; Xie, W.; Shiomi, T.; Zalk, R.; Lacampagne, A.; Marks, A.R. Ryanodine receptor oxidation causes intracellular calcium leak and muscle weakness in aging. Cell Metab. 2011, 14, 196–207. [Google Scholar] [CrossRef] [Green Version]
  217. Delrio-Lorenzo, A.; Rojo-Ruiz, J.; Alonso, M.T.; García-Sancho, J. Sarcoplasmic reticulum Ca2+ decreases with age and correlates with the decline in muscle function in Drosophila. J. Cell Sci. 2020, 133, jcs240879. [Google Scholar] [CrossRef]
  218. Qaisar, R.; Pharaoh, G.; Bhaskaran, S.; Xu, H.; Ranjit, R.; Bian, J.; Ahn, B.; Georgescu, C.; Wren, J.D.; Van Remmen, H. Restoration of sarcoplasmic reticulum Ca2+ ATPase (SERCA) activity prevents age-related muscle atrophy and weakness in mice. Int. J. Mol. Sci. 2020, 22, 37. [Google Scholar] [CrossRef]
  219. Qaisar, R.; Bhaskaran, S.; Ranjit, R.; Sataranatarajan, K.; Premkumar, P.; Huseman, K.; Van Remmen, H. Restoration of SERCA ATPase prevents oxidative stress-related muscle atrophy and weakness. Redox Biol. 2019, 20, 68–74. [Google Scholar] [CrossRef]
  220. Wu, M.M.; Buchanan, J.; Luik, R.M.; Lewis, R.S. Ca2+ store depletion causes STIM1 to accumulate in ER regions closely associated with the plasma membrane. J. Cell Biol. 2006, 174, 803–813. [Google Scholar] [CrossRef]
  221. Navarro-Borelly, L.; Somasundaram, A.; Yamashita, M.; Ren, D.; Miller, R.J.; Prakriya, M. STIM1–Orai1 interactions and Orai1 conformational changes revealed by live-cell FRET microscopy. J. Physiol. 2008, 586, 5383–5401. [Google Scholar] [CrossRef]
  222. Stiber, J.; Hawkins, A.; Zhang, Z.-S.; Wang, S.; Burch, J.; Graham, V.; Ward, C.C.; Seth, M.; Finch, E.; Malouf, N. STIM1 signalling controls store-operated calcium entry required for development and contractile function in skeletal muscle. Nat. Cell Biol. 2008, 10, 688–697. [Google Scholar] [CrossRef] [Green Version]
  223. Wei-LaPierre, L.; Carrell, E.M.; Boncompagni, S.; Protasi, F.; Dirksen, R.T. Orai1-dependent calcium entry promotes skeletal muscle growth and limits fatigue. Nat. Commun. 2013, 4, 2805. [Google Scholar] [CrossRef]
  224. Zhao, X.; Weisleder, N.; Thornton, A.; Oppong, Y.; Campbell, R.; Ma, J.; Brotto, M. Compromised store-operated Ca2+ entry in aged skeletal muscle. Aging Cell 2008, 7, 561–568. [Google Scholar] [CrossRef] [Green Version]
  225. Thornton, A.M.; Zhao, X.; Weisleder, N.; Brotto, L.S.; Bougoin, S.; Nosek, T.M.; Reid, M.; Hardin, B.; Pan, Z.; Ma, J. Store-operated Ca2+ entry (SOCE) contributes to normal skeletal muscle contractility in young but not in aged skeletal muscle. Aging (Albany NY) 2011, 3, 621. [Google Scholar] [CrossRef] [Green Version]
  226. Edwards, J.N.; Blackmore, D.G.; Gilbert, D.F.; Murphy, R.M.; Launikonis, B.S. Store-operated calcium entry remains fully functional in aged mouse skeletal muscle despite a decline in STIM1 protein expression. Aging Cell 2011, 10, 675–685. [Google Scholar] [CrossRef]
  227. Weisleder, N.; Brotto, M.; Komazaki, S.; Pan, Z.; Zhao, X.; Nosek, T.; Parness, J.; Takeshima, H.; Ma, J. Muscle aging is associated with compromised Ca2+ spark signaling and segregated intracellular Ca2+ release. J. Cell Biol. 2006, 174, 639–645. [Google Scholar] [CrossRef]
  228. Nagaraj, R.Y.; Nosek, C.M.; Brotto, M.A.; Nishi, M.; Takeshima, H.; Nosek, T.M.; Ma, J. Increased susceptibility to fatigue of slow-and fast-twitch muscles from mice lacking the MG29 gene. Physiol. Genom. 2000, 4, 43–49. [Google Scholar] [CrossRef]
  229. Peterson, M.D.; Sen, A.; Gordon, P.M. Influence of resistance exercise on lean body mass in aging adults: A meta-analysis. Med. Sci. Sport. Exerc. 2011, 43, 249. [Google Scholar] [CrossRef] [Green Version]
  230. Zeng, Z.; Liang, J.; Wu, L.; Zhang, H.; Lv, J.; Chen, N. Exercise-induced autophagy suppresses sarcopenia through Akt/mTOR and Akt/FoxO3a signal pathways and AMPK-mediated mitochondrial quality control. Front. Physiol. 2020, 11, 583478. [Google Scholar] [CrossRef]
  231. Colman, R.J.; Anderson, R.M.; Johnson, S.C.; Kastman, E.K.; Kosmatka, K.J.; Beasley, T.M.; Allison, D.B.; Cruzen, C.; Simmons, H.A.; Kemnitz, J.W. Caloric restriction delays disease onset and mortality in rhesus monkeys. Science 2009, 325, 201–204. [Google Scholar] [CrossRef] [Green Version]
  232. Colman, R.J.; Beasley, T.M.; Allison, D.B.; Weindruch, R. Attenuation of sarcopenia by dietary restriction in rhesus monkeys. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 2008, 63, 556–559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Valdez, G.; Tapia, J.C.; Kang, H.; Clemenson, G.D., Jr.; Gage, F.; Lichtman, J.W.; Sanes, J.R. Attenuation of age-related changes in mouse neuromuscular synapses by caloric restriction and exercise. Proc. Natl. Acad. Sci. USA 2010, 107, 14863–14868. [Google Scholar] [CrossRef] [PubMed]
  234. Shen, W.; Chen, J.; Zhou, J.; Martin, C.K.; Ravussin, E.; Redman, L.M. Effect of 2-year caloric restriction on organ and tissue size in nonobese 21-to 50-year-old adults in a randomized clinical trial: The CALERIE study. Am. J. Clin. Nutr. 2021, 114, 1295–1303. [Google Scholar] [CrossRef] [PubMed]
  235. Miller, S.; Wolfe, R.R. The danger of weight loss in the elderly. J. Nutr. Health Aging 2008, 12, 487–491. [Google Scholar] [CrossRef] [PubMed]
  236. Ter Borg, S.; Luiking, Y.; van Helvoort, A.; Boirie, Y.; Schols, J.; de Groot, C. Low levels of branched chain amino acids, eicosapentaenoic acid and micronutrients are associated with low muscle mass, strength and function in community-dwelling older adults. J. Nutr. Health Aging 2019, 23, 27–34. [Google Scholar] [CrossRef] [PubMed]
  237. Fuchs, C.J.; Hermans, W.J.; Holwerda, A.M.; Smeets, J.S.; Senden, J.M.; van Kranenburg, J.; Gijsen, A.P.; Wodzig, W.K.; Schierbeek, H.; Verdijk, L.B. Branched-chain amino acid and branched-chain ketoacid ingestion increases muscle protein synthesis rates in vivo in older adults: A double-blind, randomized trial. Am. J. Clin. Nutr. 2019, 110, 862–872. [Google Scholar] [CrossRef] [Green Version]
  238. McDonald, C.K.; Ankarfeldt, M.Z.; Capra, S.; Bauer, J.; Raymond, K.; Heitmann, B.L. Lean body mass change over 6 years is associated with dietary leucine intake in an older Danish population. Br. J. Nutr. 2016, 115, 1556–1562. [Google Scholar] [CrossRef] [Green Version]
  239. Devries, M.C.; McGlory, C.; Bolster, D.R.; Kamil, A.; Rahn, M.; Harkness, L.; Baker, S.K.; Phillips, S.M. Leucine, not total protein, content of a supplement is the primary determinant of muscle protein anabolic responses in healthy older women. J. Nutr. 2018, 148, 1088–1095. [Google Scholar] [CrossRef] [Green Version]
  240. Wolfe, R.R. Branched-chain amino acids and muscle protein synthesis in humans: Myth or reality? J. Int. Soc. Sport. Nutr. 2017, 14, 30. [Google Scholar] [CrossRef] [Green Version]
  241. Ko, C.-H.; Wu, S.-J.; Wang, S.-T.; Chang, Y.-F.; Chang, C.-S.; Kuan, T.-S.; Chuang, H.-Y.; Chang, C.-M.; Chou, W.; Wu, C.-H. Effects of enriched branched-chain amino acid supplementation on sarcopenia. Aging (Albany NY) 2020, 12, 15091. [Google Scholar] [CrossRef]
  242. Ryu, D.; Mouchiroud, L.; Andreux, P.A.; Katsyuba, E.; Moullan, N.; Nicolet-dit-Félix, A.A.; Williams, E.G.; Jha, P.; Lo Sasso, G.; Huzard, D. 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]
  243. Bhasin, S.; Woodhouse, L.; Casaburi, R.; Singh, A.B.; Mac, R.P.; Lee, M.; Yarasheski, K.E.; Sinha-Hikim, I.; Dzekov, C.; Dzekov, J. Older men are as responsive as young men to the anabolic effects of graded doses of testosterone on the skeletal muscle. J. Clin. Endocrinol. Metab. 2005, 90, 678–688. [Google Scholar] [CrossRef]
  244. Calof, O.M.; Singh, A.B.; Lee, M.L.; Kenny, A.M.; Urban, R.J.; Tenover, J.L.; Bhasin, S. Adverse events associated with testosterone replacement in middle-aged and older men: A meta-analysis of randomized, placebo-controlled trials. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 2005, 60, 1451–1457. [Google Scholar] [CrossRef] [Green Version]
  245. Borst, S.E.; Yarrow, J.F.; Conover, C.F.; Nseyo, U.; Meuleman, J.R.; Lipinska, J.A.; Braith, R.W.; Beck, D.T.; Martin, J.S.; Morrow, M. Musculoskeletal and prostate effects of combined testosterone and finasteride administration in older hypogonadal men: A randomized, controlled trial. Am. J. Physiol.-Endocrinol. Metab. 2014, 306, E433–E442. [Google Scholar] [CrossRef] [Green Version]
  246. Papanicolaou, D.A.; Ather, S.; Zhu, H.; Zhou, Y.; Lutkiewicz, J.; Scott, B.; Chandler, J. A phase IIA randomized, placebo-controlled clinical trial to study the efficacy and safety of the selective androgen receptor modulator (SARM), MK-0773 in female participants with sarcopenia. J. Nutr. Health Aging 2013, 17, 533–543. [Google Scholar] [CrossRef]
  247. Latres, E.; Pangilinan, J.; Miloscio, L.; Bauerlein, R.; Na, E.; Potocky, T.B.; Huang, Y.; Eckersdorff, M.; Rafique, A.; Mastaitis, J. Myostatin blockade with a fully human monoclonal antibody induces muscle hypertrophy and reverses muscle atrophy in young and aged mice. Skelet. Muscle 2015, 5, 34. [Google Scholar] [CrossRef] [Green Version]
  248. Becker, C.; Lord, S.R.; Studenski, S.A.; Warden, S.J.; Fielding, R.A.; Recknor, C.P.; Hochberg, M.C.; Ferrari, S.L.; Blain, H.; Binder, E.F. Myostatin antibody (LY2495655) in older weak fallers: A proof-of-concept, randomised, phase 2 trial. Lancet Diabetes Endocrinol. 2015, 3, 948–957. [Google Scholar] [CrossRef]
  249. Rooks, D.; Praestgaard, J.; Hariry, S.; Laurent, D.; Petricoul, O.; Perry, R.G.; Lach-Trifilieff, E.; Roubenoff, R. Treatment of sarcopenia with bimagrumab: Results from a phase II, randomized, controlled, proof-of-concept study. J. Am. Geriatr. Soc. 2017, 65, 1988–1995. [Google Scholar] [CrossRef]
  250. Alway, S.E.; McCrory, J.L.; Kearcher, K.; Vickers, A.; Frear, B.; Gilleland, D.L.; Bonner, D.E.; Thomas, J.M.; Donley, D.A.; Lively, M.W. Resveratrol enhances exercise-induced cellular and functional adaptations of skeletal muscle in older men and women. J. Gerontol. Ser. A Biomed. Sci. Med. Sci. 2017, 72, 1595–1606. [Google Scholar] [CrossRef]
Figure 1. Activation of mTORC1 by growth factors and amino acids. Activation of mTORC1 requires the integration of extracellular and intracellular signals. Binding of insulin to the insulin receptor induces a cascade of signaling events that culminate in Akt-mediated inhibition of TSC1/2. Phosphorylation of TSC1/2 by Akt prevents TSC1/2 GAP activity towards RHEB, allowing for activation of mTORC1. In response to amino acid availability, CASTOR1 and Sestrin2 inhibition of GATOR2 is relieved, allowing for GATOR2 inhibition of GATOR1. Inhibition of GATOR1 allows for Rag-mediated activation of mTORC1. Phosphorylation of p70S6K1, 4e-BP1- and eEF2 kinase by mTORC1 enhances protein translation initiation and elongation. Activated p70S6K1 can exert negative feedback on Akt via inhibition of the insulin signaling cascade.
Figure 1. Activation of mTORC1 by growth factors and amino acids. Activation of mTORC1 requires the integration of extracellular and intracellular signals. Binding of insulin to the insulin receptor induces a cascade of signaling events that culminate in Akt-mediated inhibition of TSC1/2. Phosphorylation of TSC1/2 by Akt prevents TSC1/2 GAP activity towards RHEB, allowing for activation of mTORC1. In response to amino acid availability, CASTOR1 and Sestrin2 inhibition of GATOR2 is relieved, allowing for GATOR2 inhibition of GATOR1. Inhibition of GATOR1 allows for Rag-mediated activation of mTORC1. Phosphorylation of p70S6K1, 4e-BP1- and eEF2 kinase by mTORC1 enhances protein translation initiation and elongation. Activated p70S6K1 can exert negative feedback on Akt via inhibition of the insulin signaling cascade.
Cells 12 00249 g001
Figure 2. Proposed model of mTORC1-mediated muscle atrophy in sarcopenia. Aged skeletal muscle exhibits anabolic resistance to hyperaminoacidemia characterized by an impaired protein synthetic response. Basal mTORC1 exhibits sustained activation in aged skeletal muscle, that can contribute to neuromuscular junction instability, ER stress, and impaired mitophagy. Poor cellular quality control mechanisms result in a compromised mitochondrial pool, leading to impaired bioenergetics and enhanced reactive oxygen species production.
Figure 2. Proposed model of mTORC1-mediated muscle atrophy in sarcopenia. Aged skeletal muscle exhibits anabolic resistance to hyperaminoacidemia characterized by an impaired protein synthetic response. Basal mTORC1 exhibits sustained activation in aged skeletal muscle, that can contribute to neuromuscular junction instability, ER stress, and impaired mitophagy. Poor cellular quality control mechanisms result in a compromised mitochondrial pool, leading to impaired bioenergetics and enhanced reactive oxygen species production.
Cells 12 00249 g002
Figure 3. Autophagy and mitophagy pathways. Autophagy induction is negatively regulated by mTORC1 signaling through direct inhibition of ULK1 activity. During periods of nutrient deprivation, mTORC1 activity is blunted, thereby releasing inhibition of the ULK1 initiation complex. Additionally, activation of AMPK under conditions of energetic stress directly facilitates ULK1 activation. Recruitment of the Beclin-VPS34 core complex to the ER facilitates LC3 lipidation and autophagosome membrane expansion. Loss of mitochondrial Δψm triggers OMM PINK1 accumulation and Parkin recruitment, which results in polyubiquitination and sequestration of damaged mitochondria in autophagosomes for lysosomal degradation.
Figure 3. Autophagy and mitophagy pathways. Autophagy induction is negatively regulated by mTORC1 signaling through direct inhibition of ULK1 activity. During periods of nutrient deprivation, mTORC1 activity is blunted, thereby releasing inhibition of the ULK1 initiation complex. Additionally, activation of AMPK under conditions of energetic stress directly facilitates ULK1 activation. Recruitment of the Beclin-VPS34 core complex to the ER facilitates LC3 lipidation and autophagosome membrane expansion. Loss of mitochondrial Δψm triggers OMM PINK1 accumulation and Parkin recruitment, which results in polyubiquitination and sequestration of damaged mitochondria in autophagosomes for lysosomal degradation.
Cells 12 00249 g003
Figure 4. Potential therapeutic interventions to treat sarcopenia. Calorie restriction, rapamycin, and rapalogs can enhance skeletal muscle quality through inhibition of mTORC1 signaling and enhancement of autophagy and cellular quality control mechanisms. Administration of BAM15 and urolithin A improves muscle mass and quality in sarcopenic obesity through depolarization of mitochondria and subsequent mitophagic clearance, thereby dynamically maintaining the health of the mitochondrial pool. Testosterone and SARMs improve skeletal muscle mass through androgen receptor (AR)-mediated mechanisms, while finasteride is used to diminish androgenic signaling by inhibiting dihydrotestosterone (DHT) production. Antibodies used to target both myostatin (MSTN) or the activin type II receptor inhibit myostatin-mediated signaling upstream of SMAD2/3, thereby reducing catabolic gene expression and facilitating skeletal muscle anabolism.
Figure 4. Potential therapeutic interventions to treat sarcopenia. Calorie restriction, rapamycin, and rapalogs can enhance skeletal muscle quality through inhibition of mTORC1 signaling and enhancement of autophagy and cellular quality control mechanisms. Administration of BAM15 and urolithin A improves muscle mass and quality in sarcopenic obesity through depolarization of mitochondria and subsequent mitophagic clearance, thereby dynamically maintaining the health of the mitochondrial pool. Testosterone and SARMs improve skeletal muscle mass through androgen receptor (AR)-mediated mechanisms, while finasteride is used to diminish androgenic signaling by inhibiting dihydrotestosterone (DHT) production. Antibodies used to target both myostatin (MSTN) or the activin type II receptor inhibit myostatin-mediated signaling upstream of SMAD2/3, thereby reducing catabolic gene expression and facilitating skeletal muscle anabolism.
Cells 12 00249 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Paez, H.G.; Pitzer, C.R.; Alway, S.E. Age-Related Dysfunction in Proteostasis and Cellular Quality Control in the Development of Sarcopenia. Cells 2023, 12, 249. https://doi.org/10.3390/cells12020249

AMA Style

Paez HG, Pitzer CR, Alway SE. Age-Related Dysfunction in Proteostasis and Cellular Quality Control in the Development of Sarcopenia. Cells. 2023; 12(2):249. https://doi.org/10.3390/cells12020249

Chicago/Turabian Style

Paez, Hector G., Christopher R. Pitzer, and Stephen E. Alway. 2023. "Age-Related Dysfunction in Proteostasis and Cellular Quality Control in the Development of Sarcopenia" Cells 12, no. 2: 249. https://doi.org/10.3390/cells12020249

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