Next Article in Journal
Functional Insights into the Sphingolipids C1P, S1P, and SPC in Human Fibroblast-like Synoviocytes by Proteomic Analysis
Previous Article in Journal
Transcriptomic Identification of Potential C2H2 Zinc Finger Protein Transcription Factors in Pinus massoniana in Response to Biotic and Abiotic Stresses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Epigenetics of Skeletal Muscle Atrophy

1
Department of Biochemistry, Jeonbuk National University Medical School, Jeonju 54896, Republic of Korea
2
School of Pharmacy and Institute of New Drug Development, Jeonbuk National University, Jeonju 54896, Republic of Korea
*
Author to whom correspondence should be addressed.
These authors contribute equally to this work.
Int. J. Mol. Sci. 2024, 25(15), 8362; https://doi.org/10.3390/ijms25158362
Submission received: 2 July 2024 / Revised: 23 July 2024 / Accepted: 25 July 2024 / Published: 31 July 2024
(This article belongs to the Section Molecular Genetics and Genomics)

Abstract

:
Skeletal muscle atrophy, characterized by diminished muscle strength and mass, arises from various causes, including malnutrition, aging, nerve damage, and disease-related secondary atrophy. Aging markedly escalates the prevalence of sarcopenia. Concurrently, the incidence of muscle atrophy significantly rises among patients with chronic ailments such as heart failure, diabetes, and chronic obstructive pulmonary disease (COPD). Epigenetics plays a pivotal role in skeletal muscle atrophy. Aging elevates methylation levels in the promoter regions of specific genes within muscle tissues. This aberrant methylation is similarly observed in conditions like diabetes, neurological disorders, and cardiovascular diseases. This study aims to explore the relationship between epigenetics and skeletal muscle atrophy, thereby enhancing the understanding of its pathogenesis and uncovering novel therapeutic strategies.

1. Introduction

Skeletal muscles, comprising over 600 parts and constituting approximately 40% of body weight, are integral to maintaining body temperature, facilitating physical activity, and regulating energy metabolism. Skeletal muscle atrophy primarily arises from a disparity where protein degradation surpasses synthesis, resulting in a gradual loss of muscle mass and function. Factors contributing to this condition are categorized as congenital, such as amyotrophic lateral sclerosis (ALS), spinal muscular atrophy (SMA), and Duchenne muscular dystrophy (DMD), or acquired, including chronic kidney disease (CKD), diabetes, chronic heart failure, denervation, sarcopenia, and cancer cachexia [1]. During the progression of skeletal muscle atrophy, pathological processes like inflammation, oxidative stress, autophagy, endoplasmic reticulum stress, and mitochondrial dysfunction are pivotal, either independently or collectively exacerbating the condition. This complexity complicates clinical treatment approaches. Current preventive and therapeutic measures encompass exercise, dietary supplements, and medication; nonetheless, an effective treatment remains elusive [2].
Since Conrad Waddington’s initial definition of epigenetics in the early 1940s, interest and research in this field have surged [3]. Significant advancements have been achieved in elucidating epigenetic mechanisms that regulate gene expression. Epigenetics posits that the DNA content in somatic cells remains constant despite environmental changes, with gene expression variations primarily due to chromatin state differences. Beyond chromatin structure, epigenetic inheritance influences gene expression by modulating mRNA transcription and translation. It is now understood that epigenetic modifications, through reversible covalent changes to histones and nucleic acids, regulate chromatin structure and gene expression dynamically, adapting to environmental conditions. The expanding scope of epigenetics now encompasses numerous pathological conditions, including cancer [4,5].
Epigenetic alterations in skeletal muscle manifest with aging, predominantly through increased methylation levels. The methylation status of 200 CpG sites is predictive of the muscle’s chronological age. Additionally, various muscle atrophy models exhibit modifications in histone acetylation and its regulatory factors. Histone acetylases and deacetylases modulate the expression of numerous pathways and genes crucial to skeletal muscle function. Concurrently, non-coding RNAs play a significant role in muscle atrophy [6]. This article aims to offer researchers new insights and directions by focusing on epigenetic modifications and their implications in skeletal muscle atrophy.

2. Pathogenesis of Skeletal Muscle Atrophy

Skeletal muscle atrophy is a debilitating condition lacking targeted treatments and remains poorly understood. Oxidative stress and inflammation significantly contribute to its pathogenesis. The primary sources of reactive nitrogen species and reactive oxygen species in skeletal muscle are mitochondria, xanthine oxidase, uncoupled nitric oxide synthase, and nicotinamide adenine dinucleotide phosphate oxidase [1,7]. Oxidative stress induces damage to proteins, lipids, and DNA within skeletal muscle, leading to protein degradation, cell membrane damage, and apoptosis [7]. Inflammatory factors inhibit protein synthesis and enhance proteolysis via their ligands, thereby exacerbating muscle atrophy. Inflammation also impacts other tissues, such as the digestive system, adipose tissue, and liver, indirectly affecting skeletal muscles [8]. The ubiquitin–proteasome system, autophagy–lysosome system, calpain system, and caspase system are key mechanisms driving skeletal muscle atrophy [9] (Figure 1).

2.1. Ubiquitin–Proteasome System

Muscle atrophy F-box/MAFbx (atrogin-1) and muscle ring finger-1 (MuRF-1) are key ubiquitin-E3 ligases involved in skeletal muscle atrophy. MuRF-1 predominantly ubiquitinates myosin heavy chain, myosin light chains 1 and 2, troponin I, and myosin binding protein C [9], whereas atrogin-1 targets eukaryotic translation initiation factor 3 subunit f (eIF3-f), myosin heavy chain, myogenic differentiation antigen (MyoD), and other sarcomeric proteins such as desmin and the intermediate filament protein vimentin [7,10]. Additional ubiquitin-E3 ligases, including ASB2β (associated with mitochondria, contractile proteins, protein synthesis, UPS-mediated protein degradation, and cytoskeleton/sarcomeres) [11], Mindbomb-1 (Mib1, which regulates ubiquitination and proteasome-dependent degradation of Actn3) [12], and NEDD4 (which modulates the ubiquitination of KLF15 and PDLIM7) [13,14], also play critical roles in skeletal muscle atrophy.

2.2. Autophagy–Lysosome System

In various types of skeletal muscle atrophy, autophagy flux alterations are common, characterized by an increased LC3BII/LC3B1 ratio and elevated insoluble p62 levels. Studies have demonstrated that in chronic autophagy related 7 (Atg7)−/− mice, the proportion of centralized nuclei in extensor digitorum longus (EDL) muscle increases, accompanied by decreased twitch and tetanic force, as well as heightened apoptosis [15]. Furthermore, in sepsis-induced skeletal muscle atrophy, Atg7 gene knockout (KO) exacerbates muscle atrophy, reduces blood glucose levels, and raises ketone body concentrations [16]. Conversely, Doerr et al. reported that knocking down autophagy related 5 (Atg5), essential for autophagosome formation, significantly improves soleus muscle atrophy and mitigates doxorubicin-induced muscle strength loss in female Sprague Dawley rats [17].

2.3. Caspases

Caspases are conserved cysteine proteases vital for cell death and inflammation, primarily encompassing caspases 1–12 and 14 [18]. Caspase-3, which cleaves actomyosin, plays a pivotal role in skeletal muscle atrophy. Elevated caspase activity has been observed in CKD, cancer cachexia, and disuse-induced skeletal muscle atrophy [19,20,21]. In caspase-3 KO mice, the apoptotic signaling pathway is significantly diminished, leading to notable improvements in denervation-induced muscle atrophy [22]. Additionally, knocking down caspase-9 significantly enhances the expression of genes associated with cell adhesion, proliferation, growth, development, and division regulation while inhibiting stress response and cell death-related genes [23].

2.4. Calpain

Calpains are a family of calcium-activated cysteine proteases that cleave sarcomeres in skeletal muscle, releasing actin and myosin, thus rendering them susceptible to ubiquitination and subsequent degradation [24]. Elevated calpain activity has been observed in conditions such as hydrogen peroxide (H2O2) exposure, inflammation, diabetes, cachexia, denervation, and disuse-induced skeletal muscle atrophy [25,26,27,28,29].

3. The Pathway Related to Skeletal Muscle Atrophy

The primary pathway implicated in skeletal muscle atrophy is the phosphoinositide 3-kinase (PI3K)/serine/threonine-specific protein kinase (AKT) pathway, which regulates the phosphorylation of the mechanistic target of rapamycin kinase (mTOR) and its downstream p70 ribosomal protein S6 kinase (P70S6K) and eIF4E-binding protein 1 (4EBP1), thereby influencing protein synthesis. This pathway also modulates the phosphorylation of forkhead box O (Foxo), promoting the transcription of MuRF-1 and atrogin-1 (ubiquitin-E3 ligase members) and affecting autophagy flux [30]. In patients with pancreatic carcinoma-induced cachexia, AKT protein levels are reduced by 55% compared to non-cachexia patients. Furthermore, the phosphorylation or abundance of Foxo1, Foxo3a, mTOR, and their downstream p70S6K are variably decreased [31]. In patients with sarcopenia, insulin-like growth factor 1 (IGF-1) mRNA levels decrease by 45% and AKT phosphorylation by approximately 30% [32]. Goncalves et al. found that in AKT1-KO mice, the weights of the quadriceps, EDL, and gastrocnemius muscles (predominantly type II fibers) are reduced, while AKT2-KO leads to reduced EDL and gastrocnemius muscle weights [33]. Additionally, AKT2-KO reduces AMPK phosphorylation, myocyte enhancer factor 2A expression in muscle, and mitochondrial DNA (mtDNA) abundance [34].
Pro-inflammatory cytokines such as interleukin-1 (IL-1), tumor necrosis factor-alpha (TNF-α), and interleukin-6 (IL-6) facilitate muscle proteolysis. Studies show elevated IL-6, TNF-α, and C-reactive protein (CRP) levels correlate with diminished grip strength and muscle mass [35,36]. The nuclear factor kappa-light-chain-enhancer of the activated B cell (NF-κB) pathway, related to inflammation, also significantly contributes to skeletal muscle atrophy, particularly in cancer cachexia, sarcopenia, and inflammation-induced muscle atrophy [37,38,39,40]. For instance, NF-κB pathway activity in the muscles of 70-year-olds is approximately four times higher than in young adults [40], and in patients with gastric cancer, phospho-p65 increases by approximately 25% while IκBα expression decreases by about 25% [41]. Nuclear factor kappa B subunit 1 (Nfkb1) KO notably reduces disuse-induced skeletal muscle atrophy in mice, primarily by enhancing the muscle fiber area of fast fibers [42].
The AMPK pathway is essential in skeletal muscle atrophy, as it can inhibit protein synthesis by modulating the phosphorylation of mTOR and eukaryotic initiation factor 2 [43]. Additionally, AMPK promotes mitochondrial biosynthesis and reduces oxidative stress through SIRT1 and PGC1-α [44]. In patients with non-small cell lung cancer, levels of various AMPK subunits are elevated, likely due to AMPK’s protective activation, as suggested by Raun [45]. AMPKα1/α2 KO in mice leads to reduced body weight and muscle fibers, impairing muscle function [46]. Conversely, Guo et al. found that AMPKα2 KO reduces protein degradation and mitigates denervation-induced muscle atrophy, indicating model-specific effects [47].
Myostatin, also known as growth differentiation factor-8 (GDF-8), is a member of the TGFβ family that negatively regulates skeletal muscle growth and development [48]. It induces muscle atrophy by affecting the expression of ubiquitin-E3 ligases (MuRF-1 and atrogin-1) via the Smad pathway [49]. Elevated myostatin levels have been observed in patients with gastric and lung cancers [50,51]. Higher serum myostatin levels are also found in patients with hepatocellular carcinoma compared to healthy individuals, correlating with a worse prognosis [52,53]. In the elderly with sarcopenia, muscle and serum myostatin levels are increased compared to younger individuals [54,55]. In patients with inflammatory bowel disease, serum myostatin levels correlate with sarcopenia severity [56]. Myostatin KO has been shown to improve muscle atrophy caused by glucocorticoids and aging [57,58], while its overexpression negates muscle hypertrophy induced by IGF-1 overexpression [59]. However, myostatin−/− animals, despite having greater muscle volume and body weight, exhibit reduced specific force and fatigue resistance compared to wild-type animals [60] (Figure 2).

4. Epigenetics in Different Kinds of Skeletal Muscle Atrophy

Skeletal muscle atrophy, characterized by disrupted metabolic homeostasis, often involves redox adaptation and mitochondrial dysfunction, leading to metabolic reprogramming. These metabolic disorders are intricately linked to epigenetics and encompass DNA modifications, histone modifications, chromatin remodeling, and RNA epigenetics [61,62]. This section will explore the roles of various epigenetic modifications in skeletal muscle atrophy.

4.1. DNA Modification

4.1.1. The Role of DNA Modification in Skeletal Muscle Atrophy

At least 17 chemical modifications have been identified in DNA, playing key roles in various biological processes and diseases, including development, aging, and cancer. In mammalian genomes, the most prevalent DNA modification is the methylation at the fifth carbon of cytosine (5-methylcytosine or 5mC), with 70 to 80% of CpGs being methylated. Other notable DNA modifications in mammals include 5-hydroxymethylcytosine (5hmC), 5-carboxylcytosine (5caC), and 5-formylcytosine (5fC) [63].
In age-matched patients with sarcopenia, whole blood samples revealed that the methylation levels of CTSB_15, CXCL12_22, and FGF2_30 were lower than those of the normal group, while CTSB_17 and FGF19_28 exhibited higher methylation levels. Notably, FGF2_30 methylation positively correlated with appendicular skeletal muscle index, grip strength, and gait speed [64]. A weighted sum of sarcopenia-driven CpG-site methylation levels was significantly elevated in patients with sarcopenia and negatively correlated with vastus lateralis anatomical cross-sectional area, maximum isometric elbow flexion, and knee extension [65]. Kyoto encyclopedia of genes and genomes (KEGG) analysis indicated that genes related to methylation differences in sarcopenia were mainly enriched in pathways involving actin cytoskeleton regulation, muscle function, and energy metabolism [66].
In the longissimus dorsi muscle of 7-year-old pigs, methylated DNA immunoprecipitation sequencing revealed a non-significant decrease in 5mC levels compared to half-year-old pigs. However, DNA methyltransferase (DNMT)3b mRNA levels significantly increased. A correlation analysis of differentially methylated region (DMR)-mRNA pairs identified 9234 differentially methylated regions between these groups. Hypomethylated gene body regions were significantly enriched in processes related to protein catabolism, energy metabolism, and proteolysis. Additionally, increased Foxo3 mRNA levels and decreased methylation of its gene body were observed in the skeletal muscle of 7-year-old pigs, while FGFR1, a gene inhibiting muscle atrophy, showed downregulated expression and hypermethylation [67].
In tetrodotoxin-induced skeletal muscle atrophy, murf-1 and atrogin-1 exhibited reduced DNA methylation and increased mRNA levels. Post-exercise, murf-1 DNA methylation significantly increased [68]. In the soleus muscle of mice, one week post-immobilization, the nNOS gene methylation level significantly rose [69]. Conversely, UBR5, an E3 ubiquitinase, showed reduced methylation in disused muscle atrophy, with exercise promoting its methylation level [70].

4.1.2. The Protective Role of DNMT on Skeletal Muscle Atrophy

DNA methyltransferases primarily include DNMT1, DNMT3A, and DNMT3B. DNMT3A and DNMT3B catalyze de novo DNA methylation, while DNMT1 functions as a processive DNA methyltransferase. Key DNA demethylases belong to the TET family, including TET1, TET2, and TET3. TET1 mediates the conversion of 5mC to 5hmC, with the latter two acting as intermediates in active DNA demethylation pathways [63].
In denervation-induced skeletal muscle atrophy, DNMT3A expression is significantly reduced, resulting in hypomethylation of the Gdf5 and Fn14 promoter regions. This hypomethylation impairs cell differentiation and disrupts the NF-κB and UPS, contributing to muscle atrophy [71,72].

4.2. Histone Modifications

Histones are basic proteins in the nuclei of eukaryotic cells that bind to DNA, forming nucleosomes together with 147 DNA base pairs. Each nucleosome is an octamer composed of two copies each of histones H2A, H2B, H3, and H4. Amino acid sequences of histones are conserved across species. Post-translational modifications of histones, including acylation (e.g., acetylation, butyrylation, benzoylation, glutarylation, crotonylation, lactylation), ADP-ribosylation, glycosylation, dopaminylation, methylation, serotonylation, phosphorylation, threosylation, and ubiquitination, play pivotal roles in chromatin condensation, gene transcription regulation, and DNA replication [73,74,75].

4.2.1. Histone Acetylation

Histone acetylation involves the binding of an acetyl group to a lysine residue at the N-terminus of histones. This process is mediated by “writing” enzymes known as histone acetyltransferases (HATs), “erasing” enzymes called histone deacetylases (HDACs), and “reading” enzymes containing bromodomains that recognize and bind to acetylated lysine residues. Histone acetylation disrupts protein-DNA interactions and loosens chromatin structure, thereby promoting gene transcription [76,77,78].

The Role of Histone Acetylation in Skeletal Muscle Atrophy

Histone acetylation plays a vital role in normal muscle function by regulating energy metabolism and glucose utilization in muscle cells. For instance, long-term exercise significantly increases histone H3 acetylation [79]. Ryder et al., using a peptide IP proteomic approach with an anti-acetyl-lysine antibody or a ubiquitin residue motif antibody followed by mass spectrometry, confirmed that histone acetylation increases while ubiquitination decreases in rat skeletal muscle from a cast immobilization-induced atrophy model. This change correlates with the increased transcription of genes related to muscle atrophy. Conversely, proteins associated with muscle contractions, such as alpha-actin, troponin C, myosin light chains 3 (MLC3), myosin heavy chain I, troponin T, and myosin heavy chain IIa, are deacetylated and ubiquitinated [80]. Furthermore, Kawano et al. demonstrated that in denervation-induced skeletal muscle atrophy, histone acetylation of fast and slow muscle fiber-related genes significantly decreased in EDL muscles [81]. Additionally, in aging mice, the gastrocnemius muscle weight-to-body weight ratio significantly declined in 12- and 24-month-old mice, while atrogin-1 mRNA levels increased. Acetylation of histone H3, H3K9, and H3K27 negatively correlated with age and was significantly reduced in older mice [82].

HATs Play Different Roles in Different Types of Muscular Atrophy

HATs are primarily classified into three families: GNAT, MYST, and CBP/p300. The GNAT family includes GCN5, PCAF, HAT1, Elp3, and Hpa2. The MYST family comprises Tip60 (HAT6), Sas2, Sas3, Esa1, MOF, MORF, and Hbo1. HATs facilitate DNA repair and its dissociation from the histone octamer [83].
Studies indicate that the expression of p300, Cbp, Pcaf, and Moz is elevated in immobilization, denervation, and starvation-induced muscle atrophy models, while Gcn5 expression increases only in the former two [84]. Muscle-specific knockout of Gcn5 results in increased dystrophin expression, leading to myopathy [85]. Conversely, Lee et al., demonstrated that GCN5 knockdown ameliorates starvation-induced muscle atrophy by altering p65K310 acetylation [86].
Regarding CBP/p300, activation of p300 is increased in dexamethasone, palmitic acid, or sepsis-induced skeletal muscle atrophy [87,88,89]. Knockdown of p300 improves muscle condition by modulating the acetylation of p65, FOXO1, and FOXO3a or by affecting autophagic flux [87,88,90]. Researches by Sin et al. and Liang et al. revealed that p38β MAPK phosphorylates p300 at serine-12, thereby stimulating C/EBPβ acetylation and contributing to cancer cachexia [91,92].

HDACs Play Different Roles in Congenital and Acquired Muscular Atrophy

HDACs are enzymes that remove acetyl groups from histones, regulating chromatin structure and gene expression. The HDAC family includes a large group of proteins classified into five groups: class I (HDAC 1, 2, 3, 8), class IIa (HDAC 4, 5, 7, 9), class IIb (HDAC 6, 10), class III (Sirtuins), and class IV (HDAC 11), all highly conserved across species. HDACs play critical regulatory roles in various biological processes, and their abnormal activity is associated with diseases such as neurodegenerative disorders and cancers. Research on HDACs in skeletal muscle mainly focuses on HDAC1, 4, 6, and Sirtuins [93].
Beharry et al. found that HDAC2, 4, 6, and Sirtuin1 levels significantly increased in skeletal muscle atrophy induced by denervation, immobilization, and nutrient deprivation, while HDAC1 and 3 increased in the first two models. Conversely, HDAC7 was significantly reduced in all three models [84]. Moresi et al. demonstrated that knockout of HDAC4 or HDAC5, individually or simultaneously, increased TA muscle weight compared to controls and significantly decreased muscle atrophy markers (MuRF-1 and atrogin-1) in denervation-induced atrophy model mice [94]. Additionally, HDAC4 gene KO improved autophagic flux and oxidative stress inhibition caused by denervation and promoted MHC expression by enhancing its acetylation [95]. However, in an ALS model, HDAC4 KO led to earlier ALS onset with reduced body weight, TA muscle cross-section area, and NMJ surface area compared to controls [96]. HDAC1 regulates E3 ubiquitin ligase and autophagy by modulating Foxo1a phosphorylation, and its inhibition significantly improves immobilization-induced skeletal muscle atrophy [97]. In denervation-induced muscle atrophy, HDAC6 is regulated by Foxo3a and interacts with atrogin-1 to promote MyoD ubiquitination and degradation, thus promoting muscle atrophy [98].
Several HDAC inhibitors, such as trichostatin A, butyrate, MS-275, LMK-235, NVS-HD1, tubastatin A, valproic acid, and HC toxin, have shown efficacy in various muscle wasting models (dexamethasone, sarcopenia, unloading, denervation, starvation, DMD, cancer cachexia, cigarette smoke exposure, ALS, Table 1) [87,95,97,99,100,101,102,103,104,105,106,107]. For example, in SOD1-G93A mice, a model of ALS, histone H3 acetylation was significantly reduced. Trichostatin A, an HDAC inhibitor, significantly improved histone H3 acetylation, inhibited motor neuron death and axonal degeneration, and enhanced motor function and muscle condition. The drug also ameliorated skeletal muscle atrophy caused by posterior suspension.

4.2.2. Histone Methylation

The Role of Histone Methylation in Skeletal Muscle Atrophy

Histone methylation involves the addition of methyl groups to lysine or arginine residues of histones, facilitated by histone methyltransferase (HMT). This modification can occur at various positions on lysine and arginine, resulting in monomethylation (me1), dimethylation (me2), or trimethylation (me3). Primarily targeting histones H3 and H4, histone methylation significantly influences gene expression, depending on the specific residue and methylation degree. For instance, H3K4me3, the trimethylation of lysine 4 on the H3 histone, typically occurs in promoter regions and correlates with active gene transcription. Conversely, H3K9me3, the trimethylation of lysine 9 on the H3 histone, is linked to gene silencing and heterochromatin formation. Similarly, H3K27me3, the trimethylation of lysine 27 on H3, is generally associated with gene silencing [108,109,110]. Histone methylation is involved in various pathophysiological processes, including gene expression regulation, cell differentiation and development, and the maintenance of genome stability and integrity [111]. Studies reveal that in rat gastrocnemius muscle, the H3K9me3 histone methylation modification decreases with age, showing a significant negative correlation with atrogin-1 mRNA levels and a positive correlation with relative muscle weight [82].

HMTs Protects Skeletal Muscle from Atrophy

In the advanced stage of denervation-induced skeletal muscle atrophy, the expression of protein arginine methyltransferases (PRMT) 1, CARM1 (also known as PRMT4) and PRMT5 are elevated [112]. PRMT1, an N-arginine methyltransferase, mediates the transfer of methyl groups from S-adenosyl-L-methionine (SAM or AdoMet) to the guanidine termini on arginine residues’ side chains, thus facilitating the formation of dimethylarginine [113]. This enzyme plays a key role in histone methylation and significantly impacts skeletal muscle atrophy, with research indicating that PRMT1 KO induces muscle atrophy. Additionally, PRMT1 associates with the promoter region of PRMT6, and its gene KO enhances PRMT6 expression, excessively activating Foxo3 and promoting muscle atrophy [114]. Similarly, CARM1 is pivotal in muscle atrophy, with its KO disrupting mitophagy and autophagy, leading to muscle degradation [115]. Stouth et al. demonstrated that CARM1 deficiency only reduced the exercise capacity and endurance in male mice without affecting muscle mass or denervation-induced atrophy [116].
Unlike other methyltransferases, DOT1L lacks the Su(var)3-9, enhancer of zeste, and Trithorax (SET) domain and exclusively methylates lysine [117]. Research has shown that histone methyltransferase DOT1L expression is reduced in the vastus lateralis muscle of COPD individuals with sarcopenia, and in vitro DOT1L knockdown significantly upregulates p21 expression [118].

4.2.3. γ-H2AX (a Type of Histone Phosphorylation) Is Expressed More Frequently in Skeletal Muscle Atrophy

Nearly all types of histones undergo phosphorylation at specific residues, a modification essential for transcriptional regulation, cell division, DNA damage repair, hetero-chromatin formation, and gene silencing. For instance, the phosphorylation of the third serine residue on H3 histone (H3S10) is associated with chromosome condensation, essential for chromosome segregation during cell division. Notably, DNA damage repair is significantly marked by the phosphorylation of H2A histone family member X (H2AX) at serine 139, a key indicator of aging [119,120].
Histone phosphorylation, especially of γ-H2AX, also plays a vital role in skeletal muscle atrophy. Studies have shown that in primary human muscle progenitor cells isolated from the lateral thigh and cultured for replication and senescence, γ-H2AX expression markedly increases compared to normal cells. This increase is accompanied by a significant reduction in the expression of cell differentiation markers (Myog, MyHC, MyoD), reduced fusion, decreased myotube area, and elevated TGF-β levels [121]. Additionally, skeletal muscle atrophy associated with chronic kidney disease (CKD) and spinal muscular atrophy (SMA) is linked to heightened γ-H2AX expression [122,123]. During SYUIQ-5-induced cellular senescence, γ-H2AX expression shows a negative correlation with MyHC expression and a positive correlation with muscle atrophy markers (murf-1 and atrogin-1) [124].

4.2.4. Histone Ubiquitination Reduced during Skeletal Muscle Atrophy

Histone ubiquitination is a key post-transcriptional modification process wherein small ubiquitin proteins are covalently attached to specific amino acid residues of histones. This modification can alter chromatin structure, thereby influencing transcription factor binding and gene activity. As a reversible process, ubiquitination allows for the removal of ubiquitin molecules at various intervals. Ubiquitin can attach to multiple lysine residues on histones, forming polyubiquitin chains. The varying lengths (monoubiquitination, polyubiquitination) and types of ubiquitin chains convey distinct biological information and correspond to different degradation pathways and functional outcomes. Histone ubiquitination is implicated in numerous biological processes, such as protein degradation, DNA repair, cell cycle regulation, gene silencing, and chromatin remodeling [125]. In a rat model of immobilization-induced skeletal muscle atrophy, histone ubiquitination was significantly reduced, correlating with a marked decrease in soleus muscle mass compared to the control group, and an increase in the ubiquitination of muscle contraction proteins [80].

4.2.5. Histone Lactylation May Play a Critical Role in Skeletal Muscle Atrophy

Histone lactylation is an epigenetic modification wherein lactate molecules are added to specific amino acid residues of histones, playing a significant role in cell metabolism and gene expression regulation. This modification predominantly occurs on the lysine residues of histones H3 and H4 and is mediated by enzymes such as lactation modifying enzyme (LME). Lactate, serving as a precursor for histone lysine lactylation (Kla), stimulates gene transcription within chromatin. Histone lactylation can influence the structure and function of histones, thus altering chromatin architecture and gene expression. This modification is implicated in various biological processes, including cell proliferation, energy metabolism, differentiation, and immune response. As the primary product of glycolysis, particularly under hypoxic conditions (e.g., during intense muscle exercise), lactic acid rapidly supplies energy to muscles [126]. Additionally, lactic acid functions as a vital signaling molecule, impacting several pathological and physiological processes such as intracellular reactive oxygen species, mitochondrial biogenesis, and fatty acid oxidation. It is currently posited that P300, HDAC1-3, and sirtuin1-3 enzymes can establish and remove histone lactate [127,128]. These genes have demonstrated significant roles in various models of skeletal muscle atrophy [62,88,129,130,131].

4.3. RNA Modification

RNA modification involves chemical alterations to RNA molecules post-transcription, significantly impacting their function and stability. This process is both complex and diverse, encompassing over 170 types of chemical modifications beyond the well-known m6A modification. These modifications are prevalent in both noncoding RNA and messenger RNA, playing pivotal roles in RNA stability, translation efficiency, and intracellular localization. As a key component of epigenetics, RNA modification intricately regulates gene expression by finely tuning RNA structure and function. Together, these chemical modifications form a sophisticated regulatory network essential for maintaining cellular physiological functions and contributing to the onset and progression of various diseases [132,133].

4.3.1. RNA Methylation

The Role of RNA Methylation in Skeletal Muscle Atrophy

In 1974, methylation was first identified in mRNA, with the modified adenine known as m6A being the most common post-transcriptional modification in mRNA. Typically, 1–2 m6A residues are found per 1000 nucleotides. RNA methylation plays a critical role in gene expression regulation, with N6-adenosine methylation (m6A) being the predominant chemical modification on mRNA. The m6A modification involves the addition of a methyl group at the N6 position on the adenosine residue of mRNA molecules, catalyzed by specific enzymes. This chemical modification is widespread in eukaryotes and significantly impacts mRNA processing, stability, and translation. M6A modification regulates mRNA precursor splicing and RNA nuclear export and influences mRNA translation and stability. It also promotes circular RNA translation and contributes to cell differentiation and tumor development. The dynamic and reversible nature of m6A modification involves methyltransferases (writers), demethylases (erasers), and methyl-binding proteins (readers). Methyltransferases such as methyltransferase like (METTL) 3 and METTL14 catalyze the m6A modification, while demethylases like fat mass and obesity-associated (FTO) and alpha-ketoglutarate-dependent dioxygenase alkB homolog 5 (ALKBH5) remove these methyl groups, allowing for flexible gene expression regulation [134,135].
M6A methylation of mRNA has also been observed in the skeletal muscles of various species, potentially influencing skeletal muscle differentiation. During prenatal porcine skeletal muscle development, m6A predominantly accumulates in the coding sequence and three prime untranslated region (3′-UTR), enriched in processes related to RNA binding, nucleocytoplasmic transport, and macromolecular metabolism. The m6A and mRNA levels of certain genes involved in skeletal muscle differentiation, such as MYH2 (which encodes MyHC), are significantly elevated [136]. Additionally, Gheller et al. found that m6A levels increase during the early stages of skeletal muscle regeneration but decrease during C2C12 differentiation, with m6A primarily concentrated in the coding region and 3′-UTR [137].

The Role of Methyltransferases and Demethylases in Skeletal Muscle Atrophy

Knocking down the methyltransferase METTL14 in C2C12 cells inhibits differentiation, as evidenced by decreased levels of differentiation markers MyHC, MyoD, and MyoG while promoting cell growth. Conversely, METTL3, another methyltransferase, is reduced in myoblasts. Unlike METTL14, METTL3 knockdown enhances skeletal muscle cell differentiation and inhibits proliferation [137].
In denervation-induced skeletal muscle atrophy, numerous m6A-modified RNAs are predominantly downregulated. These modifications are enriched in functions such as ubiquitin–protein ligase activity, ATP binding, zinc ion binding, transcription coactivator activity, protein phosphatase binding, RNA polymerase II cis-regulatory region sequence-specific DNA binding, DNA binding, glucocorticoid receptor binding, nucleic acid binding, and mannosyl-oligosaccharide 1,2-alpha-mannosidase activity. During this process, demethylases like FTO and ALKBH5 significantly increase. Furthermore, 3-Dezidenosine (DAA) in mouse muscle promotes ALKBH5 and FTO expression, substantially reducing m6A levels in muscles and leading to decreased muscle weight and cross-sectional area. Local injection of R-2-HydroxyglutaraTe (R-2HG) in denervation-induced skeletal muscle atrophy inhibits FTO and ALKBH5 expression, promotes m6A modification, and ameliorates muscle atrophy [138]. Liu et al. also confirmed that m6A levels decrease while ALKBH5 expression increases in the gastrocnemius muscle of denervation-induced atrophy models, with ALKBH5 targeting HDAC4 to enhance Foxo3 expression and muscle atrophy [139].

4.3.2. Other Kinds of RNA Modification

In addition to RNA methylation, other types of RNA modifications include ac4C, 2-thiouridine (s2U), A-I (adenosine to inosine) RNA editing, pseudouridine (Ψ), and the 5′-cap (m7GpppN). A-I RNA editing is a post-transcriptional modification process where adenosine is converted into inosine, catalyzed by specific adenosine deaminases acting on RNA (ADARs). This modification can alter RNA structure and function by changing encoded amino acids, affecting splicing sites, or regulating RNA stability and translation efficiency. A-I RNA editing is critically involved in various biological processes, including neuronal development, immune response, and viral infection. Defective RNA editing has been linked to numerous diseases, such as neurodegenerative disorders, cancer, and autoimmune diseases [133].
ADAR2 is an enzyme responsible for A-to-I editing in RNA molecules, highly expressed in the mammalian nervous system and significantly influencing neuronal function and gene regulation. ADAR2’s functions include regulating A-to-I RNA editing, RNA splicing, and affecting RNA stability and translation efficiency. Research has shown that ADAR2 KO can modulate high-fat diet-induced skeletal muscle atrophy via the AKT/Foxo1 pathway and inflammation. Specifically, ADAR2 gene KO promotes the phosphorylation of AKT and Foxo1 while inhibiting M1 and M2 macrophage markers (iNOS, IL-12, Arg1, Fizz1, Ym1, and IL-10) [140].

4.4. Noncoding RNA

Noncoding RNA (ncRNA) refers to RNA molecules that do not encode proteins but regulate gene expression and other biological processes. NcRNA is classified by length and function into small noncoding RNA (<200 nt) and long noncoding RNA (lncRNA) (>200 nt). Small noncoding RNA includes: (1) microRNA (miRNA): approximately 22 nucleotides long, miRNA regulates gene expression by binding to the 3′ untranslated region (3′UTR) of mRNA, inducing mRNA degradation or inhibiting translation. miRNA plays a critical role in development, cell proliferation, differentiation, and apoptosis; (2) small interfering RNA (siRNA): ranging from 21–23 nucleotides in length, siRNA uses the RNA interference (RNAi) pathway to perfectly complement target mRNA, causing its degradation and thus inhibiting gene expression. siRNA is primarily used to study gene function and treat specific diseases; (3) PIWI-interacting RNA (piRNA): spanning 24–31 nucleotides, piRNA is mainly expressed in germ cells and, by binding to Piwi proteins, inhibits transposon activity, protecting genome integrity. LncRNA has multiple regulatory functions, including gene transcription regulation, RNA splicing, chromatin structure modification, acting as molecular scaffolds, and modulating protein-RNA interactions. Small nucleolar RNA (snoRNA) primarily functions within the nucleus. Circular RNA (circRNA) is a closed circular RNA molecule found in eukaryotic cells, which can regulate gene expression at the transcriptional and post-transcriptional levels, such as acting as sponges for miRNA and RNA-binding proteins [141,142].

4.4.1. miRNA

miRNA is a small non-coding RNA molecule, approximately 22 nucleotides in length, that regulates gene expression by complementing the 3′ untranslated region (3′UTR) of target mRNA. miRNA is pivotal in various biological processes, including cell proliferation, differentiation, apoptosis, and development. The biosynthesis of miRNA involves several steps: transcription (producing primary miRNA (pri-miRNA) through RNA polymerase II transcription), splicing (processing precursor miRNA of about 70 nucleotides by the Drosha-DGCR8 complex), nuclear export (transporting from the nucleus to the cytoplasm via Exportin-5), further processing (Dicer enzyme cleaves pre-miRNA into double-stranded RNA, selecting one strand as mature miRNA (guide strand) while the other (passenger strand) is typically degraded), and forming RISC complexes (binding with Argonaute protein to create RNA-induced silencing complexes (RISC)). The RISC complex mediates gene silencing by guiding miRNA to bind to target mRNA. Various miRNAs have been shown to play vital roles in different types of skeletal muscle atrophy.
Gagan et al. demonstrated that MyoD can bind to the miR-378 gene, leading to transcriptional activation and chromatin remodeling, thus promoting miR-378 expression. MiR-378 further enhances MyoD expression by inhibiting MyoR [143]. Additionally, miR-494-3p can influence the expression of P300 and its downstream targets, MYOD and MYH2 [144].
In denervation-induced skeletal muscle atrophy, miR-142a-5p exacerbates the condition by binding to the 3′UTR of MFN1 mRNA, inhibiting its expression [145]. Overexpression of miR-29 and miR-125b-5p can ameliorate denervation and fasting-induced skeletal muscle atrophy by targeting murf-1 and TRAF6, respectively [146]. Conversely, overexpression of miR-322 and miR-542 in skeletal muscle significantly contributes to atrophy by binding to the 3′UTR of eIF4B, eIF2B5, eIF4E, eIF4G1, and eIF3M, inhibiting their expression [147].
Studies have demonstrated that aging-induced atrophic myotubes and macrophages release exosomes enriched with miR-690, which inhibits the differentiation of C2C12 and satellite cells by regulating Mef2a, Mef2c, and Mef2d expression [148]. Overexpression of miR-181a enhances mitochondrial function in aged mice by downregulating p62, Park2, and DJ-1 [149]. Soriano et al. reported that miR-181a decreases the diameter of C2C12 myotubes by inhibiting sirtuin1 expression [150]. In aged mice, the overexpression of miR-434-3p in myotubes reduces the activation of caspase-8, caspase-3, and caspase-9 by inhibiting eIF5A1 [151]. In aged rats, miR-29 plays distinct roles in kidney disease and diabetes-induced skeletal muscle atrophy. Hu et al. confirmed a significant increase in miR-29 expression in aged mice and rats, potentially mediated by Wnt-3a. MiR-29 inhibits the translation of myoblast proliferation mediators by binding to the 3′UTR of IGF-1, p85a, and B-myb, promoting aging-related indicators and leading to skeletal muscle atrophy [152]. In a zebrafish sarcopenia model, miR-128 expression is elevated, and aerobic exercise inhibits its expression, likely due to miR-128 targeting the 3′UTR of IGF-1 [153]. The NF-κB subunit p50 binds to the promoter region of miR-532-3p, and miR-532-3p binds to the 3′UTR of BAK1. Therefore, in skeletal muscle atrophy, p50 inhibits miR-532-3p expression, thereby promoting BAK1 expression and cell apoptosis [154].
In dexamethasone-induced skeletal muscle atrophy, overexpression of miR-486 and miR-182 ameliorates atrophy by reducing protein levels of Foxo1 and Foxo3, respectively [155,156]. MiR-322 targets 3′UTR inhibitors of IGF-1R and INSR, exacerbating dexamethasone-induced muscle atrophy [157]. MiR-320 binds to the 3′UTR of growth factor receptor-binding protein-2 (Grb2), mitigating dexamethasone-induced muscle atrophy [158]. Inhibition of miR-29b ameliorates dexamethasone-, denervation-, H2O2-, and TNF-α-induced muscle atrophy by regulating IGF-1 and PI3K (p85α) expression [159]. MiR-23a binds to the 3′UTR of murf-1 and atrogin-1, improving dexamethasone-induced muscle atrophy [160]. MiR-27b-3p ameliorates dexamethasone-induced atrophy by inhibiting Cbl-b expression [161].
MiR-29c ameliorates LLC-induced skeletal muscle atrophy by targeting LIF [162]. Tunicamycin-induced oral squamous cell carcinoma (OSCC) cell-conditioned medium contains exosomes enriched with miR-181a-3p, which induces skeletal muscle atrophy [163]. MiR-18a binds to the 3′UTR of IGF1, inhibiting its expression and leading to skeletal muscle atrophy [164]. Exosomes derived from mesenchymal stem cells are rich in miR-145-5p, which can improve busulfan and cyclophosphamide-induced skeletal muscle atrophy by targeting the 3′UTR of ACVR2A and ACVR1B [165].
In unilateral ureteral obstruction mouse models or CKD mice, the expression of miR-29 and miR-26a is decreased [166,167,168,169]. Overexpression of miR-27 and miR-23 ameliorates CKD-induced skeletal muscle atrophy by negatively regulating PTEN, caspase-7, and FoxO1 [19]. Additionally, miR-27a binds to myostatin [170]. Downregulation of miR-29 increases YY1 expression, leading to skeletal muscle atrophy [167]. Conversely, overexpression of miR-29b in C2C12 significantly promotes the expression of muscle atrophy-related genes and reduces muscle fiber diameter [159]. Local intramuscular injection of miR-29-loaded exosomes improves skeletal muscle atrophy and renal fibrosis in mouse models [169]. Overexpression of miR-486 in CKD mice ameliorates skeletal muscle atrophy by affecting FoxO1 and PTEN [171].
In the diabetes-induced skeletal muscle atrophy model, the expression of miR-182 and miR-23a is reduced [156,172]. Local injection of AAV-miR-23a/27a into the tibialis anterior muscle promotes phosphorylation of AKT and SMAD2/3 and inhibits myostatin expression, thus improving diabetes-induced skeletal muscle atrophy [173]. Studies have shown that miR-193b expression is increased in the serum of patients with type 2 diabetes. Inhibition of miR-193b improves muscle loss and dysfunction in db/db mice by targeting PDK1 and promoting the AKT/mTOR pathway [174].
SMN binds to the upstream genomic regions of miR-206 and miR-1 [175]. A tail vein injection of AAV-miR-298 effectively ameliorates the phenotype of mice with bulbar and SMA [176]. Following spinal cord injury, the expression of miR-499-5p and miR-208b decreases, with miR-208b targeting and binding to myostatin to reduce its expression [177].
In skeletal muscle atrophy induced by hindlimb suspension, the expression of miR-203a-3p, miR-499, and miR-6516 precursors is significantly downregulated, while the miR-206 precursor is increased [178,179,180]. Among these, miR-499 regulates MyHC expression by binding to the 3′UTR of Sox6, leading to skeletal muscle atrophy [179]. Local injection of miR-6516 into the TA muscle improves skeletal muscle atrophy caused by hindlimb immobilization [180].
Under inflammatory conditions, miR-140 negatively regulates the Wnt pathway, HEY1, and Notch 1, resulting in skeletal muscle atrophy [181]. However, Shin et al., demonstrated that knocking out the miR-140 gene does not exacerbate LPS-induced skeletal muscle atrophy [182]. Inhibiting miR-21 significantly improves myotube diameter in myosatellite cells [183]. Overexpression of miR-497-5p in C2C12 cells induces muscle atrophy by regulating the insulin signaling pathway [184]; while miR-1290 promotes skeletal muscle differentiation by targeting Foxo3a and inhibiting the expression of murf-1 and atrogin-1 [185].
Compared to controls, the expression of miR-424-5p, miR-206, miR-1, miR-27a, miR-542-3p/5p, and miR-145-5p is upregulated in muscle biopsies of patients with COPD [186,187,188,189,190]. However, miR-1 is significantly decreased in the quadriceps of patients with COPD [191]. MiR-206 expression is elevated in the muscles of G93A-SOD1 transgenic mice, SMAII mice, and denervation-induced skeletal muscle atrophy model mice, while miR-133a expression is decreased in G93A-SOD1 transgenic mice [192]. In the gastrocnemius muscle of mdx4cv mice, there is a significant decrease in miR-30 family expression [193]. MiR-1/206/133 induces neuromuscular symptoms and skeletal muscle atrophy in mice by binding to the 3′UTR of CRK [194].

4.4.2. LncRNA

LncRNA is a type of RNA molecule longer than 200 nucleotides that does not encode proteins. Despite not encoding proteins directly, lncRNA plays pivotal regulatory roles in various biological processes, including gene expression regulation, chromatin structure remodeling, RNA processing, and protein interaction. LncRNA is involved in cell differentiation and development, gene imprinting regulation, and numerous disease processes.
In skeletal muscle, lnc-MYH regulates the composition of the INO80 chromatin remodeling complex. Knockout of lnc-MYH modulates the proliferation of MuSCs cells via INO80, resulting in muscle hypertrophy [195]. During muscle satellite cell differentiation, lncMREF interacts with Smarca5 to enhance chromatin accessibility and the binding of the p300/CBP/H3K27ac complex, thereby promoting MyoD expression [196]. Cabianca et al. demonstrated that in facioscapulohumeral muscular dystrophy, the elevated expression of DBE-T (a lncRNA) induces FSHD by recruiting Ash1L (a trithorax group protein) to the FSHD locus, driving histone H3 lysine 36 dimethylation, chromatin remodeling, and 4q35 gene transcription [197]. Additionally, ChRO1, a lncRNA expressed during myoblast differentiation, influences DNA methylation and chromatin compaction in pericentromeric/centromeric regions [198]. Dum binds to the Dppa2 promoter region and recruits Dnmts, promoting DNA methylation of the Dppa2 promoter [199]. DNA methylation also affects lncRNA expression; the methylation level of the GTL2 gene increases with age [200], while ALKBH5 regulates the m6A demethylation of MYH1G-AS [201].
Compared to the control group, the expression of PRKG1-AS1, lnc-MALAT1, and Chronos is upregulated, whereas GPRC5D-AS1 and AC103740.1 are downregulated in the muscles of older individuals [202,203,204]. GPRC5D-AS1, acting as a competing endogenous RNAs (ceRNA) of miR-520d-5p, inhibits the regulation of muscle regulatory factors and cell apoptosis [205]. Inhibition of Chronos leads to muscle hypertrophy through the regulation of Bmp7 [204].
In dexamethasone-induced skeletal muscle atrophy, the expression of Atrolnc-1, Dum, MAR1, lnc-MD1, and SYISL increases, while Myolinc decreases [206,207]. Knocking out SYISL alleviates dexamethasone-induced skeletal muscle atrophy. As a sponge for miR-23a-3p, miR-103-3p, and miR-205-5p, SYISL upregulates the expression of FoxO3a, atrogin-1, and murf-1, thereby promoting skeletal muscle atrophy [207]. Lnc-IRS1, acting as the ceRNA of miR-15a, miR-15b-5p, and miR-15c-5p, regulates IRS1 expression and mitigates dexamethasone-induced skeletal muscle atrophy [208].
In multiple skeletal muscle atrophy models, the expression of lnc-MAAT, Gm4544, lnc-Myh, Gm49794, Oip5os1, and Gm46085 is reduced [209,210]. LncMAAT negatively regulates miR-29b transcription through a trans-regulatory module involving SOX6, and increases the expression of neighboring gene Mbnl1 through a cis-regulatory module, thereby improving angiotensin-induced skeletal muscle atrophy [209]. Lnc-MAR1 acts as a sponge for microRNA-487b, promoting the expression of wnt5 and improving skeletal muscle atrophy [211]. In tail suspension-treated mouse TA muscles, lnc-MD1, lnc-Myod, and IG-DMR are increased, while Gtl2 expression is reduced. In casting-induced skeletal muscle atrophy, lnc-Myod and linc-MD1 expression increases, while Dum decreases. In cancerous muscle atrophy, Gtl2, DRR, and IG-DMR expression is reduced. In starvation-induced skeletal muscle atrophy, the expression of lnc-Myod, Gtl2, DRR, Dum, H19, and IG-DMR is reduced [212].
In the gastrocnemius muscle of mice with suspension-induced skeletal muscle atrophy, the expression of lnc-MUMA and lnc-H19 is significantly reduced [213,214]. Lnc-MUMA acts as a miR-762 sponge to regulate Myog expression, thereby improving hindlimb suspension-induced skeletal muscle atrophy [213].
Compared to the control group, the expression of Atrolnc-1, lnc-MD1, Myolinc, lnc-MyoD, Dum, and MAR1 significantly increases in starved tissues or cells, while lnc-mg is reduced. Atrolnc-1 expression also increases in the muscles of cancer cachexia and CKD mice. C/EBP-alpha binds to the promoter of Atrolnc-1, promoting its expression. Overexpression of Atrolnc-1 enhances murf-1 expression induced by the NF-κB pathway through ABIN-1. Knocking down Atrolnc-1 improves skeletal muscle atrophy caused by CKD [215].
Lnc-Myh is downregulated in ALS and denervation-induced skeletal muscle atrophy, whereas H19, Dancr, Gas5, Igf2os, MiR22hg, Airn, Neat1, Snhg1, and Pvt1 are upregulated [216]. Lnc-mg promotes MuSC differentiation by acting as a competitive endogenous RNA (ceRNA) for microRNA-125b to regulate IGF-2 expression [217]. In denervation-induced skeletal muscle atrophy, Myoparr, linc-MD1, LncMyod, and H19 increase while DRR, DUM, Gtl2, and IG-DMR decrease [212]. Lnc-Myoparr interacts with Ddx17 and modulates the association between histone acetyltransferase PCAF and Ddx17 [218]. Additionally, Myoparr regulates the BMP pathway, and its knockdown ameliorates denervation-induced skeletal muscle atrophy [219].
In the gastrocnemius muscle of db/db mice, the expression of 1700047G03Rik and Gm31814 increases, while Gm20743, Gm35438, Gm36131, and A330074k22rik decrease. In palmitic acid-induced C2C12 cells, the expression patterns of the remaining five lncRNAs mirror those in the gastrocnemius muscle of db/db mice, except for the decreased expression of Gm38141 [220].

4.4.3. circRNA

Circular RNA (circRNA) is a type of non-coding RNA molecule characterized by a closed circular structure, lacking 5′ and 3′ ends. This unique structure confers high stability and resistance to exonuclease degradation. CircRNA can be classified into ex-on-circRNA (ecircRNA), exon-intron-circRNA (EIcircRNA), and intron-circRNA (ciRNA). CircRNA serves as a miRNA sponge, regulating the expression, transcription, translation, protein function, and localization of target genes.
CircTmeff1 is highly expressed in various types of skeletal muscle atrophy, including denervation-, mobilization-, dexamethasone-, TNF-α-, and angiotensin II-induced models. Overexpression of circTmeff1 activates the cGAS/STING pathway via TDP-43 and encodes TMEFF1-339aa, leading to skeletal muscle atrophy. Inhibition of circT-meff1 ameliorates skeletal muscle atrophy induced by dexamethasone, denervation, and mobilization [221]. CircCCDC91 adsorbs miR-15a, miR-15b-5p, and miR-15c-5p, activating the IGF-1-PI3K/AKT signaling pathway and regulating IRS1 expression [222].

5. Conclusions and Prospects

With the advancement of research, skeletal muscle atrophy has transitioned from a debilitating condition to one that can be delayed or even potentially cured. Currently, skeletal muscle atrophy is primarily managed through exercise, dietary supplements, and drug treatments, but targeted therapies remain elusive. Thus, improving skeletal muscle atrophy remains a significant challenge for clinicians.
Epigenetics has garnered widespread attention and achieved significant progress in recent years. This article summarizes the role of epigenetics in skeletal muscle atrophy. DNA methylation, histone modification, RNA modification, and non-coding RNA are crucial in the progression of skeletal muscle atrophy. However, the study of epigenetics in skeletal muscle atrophy is still in its infancy. Further research on molecular targets is essential, and developing viable treatments for these newly discovered processes and their associated signaling networks is imperative.

Author Contributions

All authors participated in the preparation of this article. J.D. and Q.W. conducted the literature search and wrote the original draft preparation; E.J.B. contributed to the design with the revision of this review article; E.J.B. contributed to the design with the revision of this review article and funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported, and the APC was funded by grants from the Bio & Medical Technology Development Program (2022M3E5F2017607), which was funded by the Korean government.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

Special thanks to the Bio & Medical Technology Development Program of the Korean government for providing us with the research condition.

Conflicts of Interest

The authors state that publication of this work does not involve any conflicts of interest.

Abbreviations

COPDchronic obstructive pulmonary disease
ALSamyotrophic lateral sclerosis
SMAspinal muscular atrophy
DMDDuchenne muscular dystrophy
CKDchronic kidney disease
atrogin-1muscle atrophy F-box/MAFbx
murf-1muscle ring finger-1
eIF3-feukaryotic translation initiation factor 3 subunit f
MyoDmyogenic differentiation antigen
Mib1mindbomb-1
UPSubiquitin–proteasome system
EDLextensor digitorum longus
Atg7autophagy related 7
KOknockout
Atg5autophagy related 5
H2O2hydrogen peroxide
PI3Kphosphoinositide 3-kinase
AKTserine/threonine-specific protein kinase
mTORmechanistic target of rapamycin kinase
P70S6Kp70 ribosomal protein S6 kinase
4EBP1e IF4E-binding protein 1
Foxoforkhead box O
IGF-1insulin-like growth factor 1
AMPKAMP-activated protein kinase
mtDNAmitochondrial DNA
IL-1interleukin-1
TNF-αtumor necrosis factor alpha
IL-6interleukin-6
CRPC-reactive protein
NF-κBnuclear factor kappa-light-chain-enhancer of activated B cells
Nfkb1nuclear factor kappa B subunit 1
GDF-8growth differentiation factor-8
TGFβtransforming growth factor β
5mC5-methylcytosine
DNMTDNA methyltransferase
5hmC5-hydroxymethylcytosine
5caC5-carboxylcytosine
5fC5-formylcytosine
KEGGkyoto encyclopedia of genes and genomes
DMRdifferentially methylated region
DNMTDNA methyltransferases
TETten-eleven translocation
HATshistone acetyltransferases
HDACshistone deacetylases
MLC3myosin light chains 3
HMThistone methyltransferase
me1monomethylation
me2dimethylation
me3trimethylation
PRMTprotein arginine methyltransferases
H2AXH2A histone family member X
m6AN6 adenylate methylation
METTLmethyltransferase like
FTOfat mass and obesity-associated
ALKBH5alpha-ketoglutarate-dependent dioxygenase alkB homolog 5
DAA3-Dezidenosine
R-2HGR-2-HydroxyglutaTa
s2U2-thiouridine
Ψpseudouridine
ADARsadenosine deaminases acting on RNA
ncRNAnoncoding RNA
lncRNAlong noncoding RNA
miRNAmicroRNA
3′UTR3′ untranslated region
siRNAsmall interfering RNA
RNAiRNA interference
piRNAPIWI-interacting RNA
snoRNAsmall nucleolar RNA
OSCCoral squamous cell carcinoma
ceRNAcompeting endogenous RNAs

References

  1. Zhang, H.; Qi, G.; Wang, K.; Yang, J.; Shen, Y.; Yang, X.; Chen, X.; Yao, X.; Gu, X.; Qi, L.; et al. Oxidative stress: Roles in skeletal muscle atrophy. Biochem. Pharmacol. 2023, 214, 115664. [Google Scholar] [CrossRef]
  2. Sartori, R.; Romanello, V.; Sandri, M. Mechanisms of muscle atrophy and hypertrophy: Implications in health and disease. Nat. Commun. 2021, 12, 330. [Google Scholar] [CrossRef]
  3. Deans, C.; Maggert, K.A. What do you mean, “epigenetic”? Genetics 2015, 199, 887–896. [Google Scholar] [CrossRef]
  4. Kim, U.; Lee, D.S. Epigenetic Regulations in Mammalian Cells: Roles and Profiling Techniques. Mol. Cells 2023, 46, 86–98. [Google Scholar] [CrossRef]
  5. Hogg, S.J.; Beavis, P.A.; Dawson, M.A.; Johnstone, R.W. Targeting the epigenetic regulation of antitumour immunity. Nat. Rev. Drug Discov. 2020, 19, 776–800. [Google Scholar] [CrossRef]
  6. Granic, A.; Suetterlin, K.; Shavlakadze, T.; Grounds, M.D.; Sayer, A.A. Hallmarks of ageing in human skeletal muscle and implications for understanding the pathophysiology of sarcopenia in women and men. Clin. Sci. 2023, 137, 1721–1751. [Google Scholar]
  7. Agrawal, S.; Chakole, S.; Shetty, N.; Prasad, R.; Lohakare, T.; Wanjari, M. Exploring the Role of Oxidative Stress in Skeletal Muscle Atrophy: Mechanisms and Implications. Cureus 2023, 15, e42178. [Google Scholar]
  8. Ji, Y.; Li, M.; Chang, M.; Liu, R.; Qiu, J.; Wang, K.; Deng, C.; Shen, Y.; Zhu, J.; Wang, W.; et al. Inflammation: Roles in Skeletal Muscle Atrophy. Antioxidants 2022, 11, 1686. [Google Scholar] [CrossRef]
  9. Yin, L.; Li, N.; Jia, W.; Wang, N.; Liang, M.; Yang, X.; Du, G. Skeletal muscle atrophy: From mechanisms to treatments. Pharmacol. Res. 2021, 172, 105807. [Google Scholar] [CrossRef]
  10. Bodine, S.C.; Baehr, L.M. Skeletal muscle atrophy and the E3 ubiquitin ligases MuRF1 and MAFbx/atrogin-1. Am. J. Physiol. Endocrinol. Metab. 2014, 307, E469–E484. [Google Scholar] [CrossRef]
  11. Goodman, C.A.; Davey, J.R.; Hagg, A.; Parker, B.L.; Gregorevic, P. Dynamic Changes to the Skeletal Muscle Proteome and Ubiquitinome Induced by the E3 Ligase, ASB2beta. Mol. Cell. Proteom. 2021, 20, 100050. [Google Scholar] [CrossRef]
  12. Seo, J.Y.; Kang, J.S.; Kim, Y.L.; Jo, Y.W.; Kim, J.H.; Hann, S.H.; Park, J.; Park, I.; Park, H.; Yoo, K.; et al. Maintenance of type 2 glycolytic myofibers with age by Mib1-Actn3 axis. Nat. Commun. 2021, 12, 1294. [Google Scholar] [CrossRef]
  13. D’Cruz, R.; Plant, P.J.; Pablo, L.A.; Lin, S.; Chackowicz, J.; Correa, J.; Bain, J.; Batt, J. PDLIM7 is a novel target of the ubiquitin ligase Nedd4-1 in skeletal muscle. Biochem. J. 2016, 473, 267–276. [Google Scholar] [CrossRef] [PubMed]
  14. Cid-Diaz, T.; Leal-Lopez, S.; Fernandez-Barreiro, F.; Gonzalez-Sanchez, J.; Santos-Zas, I.; Andrade-Bulos, L.J.; Rodriguez-Fuentes, M.E.; Mosteiro, C.S.; Mouly, V.; Casabiell, X.; et al. Obestatin signalling counteracts glucocorticoid-induced skeletal muscle atrophy via NEDD4/KLF15 axis. J. Cachexia Sarcopenia Muscle 2021, 12, 493–505. [Google Scholar] [CrossRef]
  15. Pare, M.F.; Baechler, B.L.; Fajardo, V.A.; Earl, E.; Wong, E.; Campbell, T.L.; Tupling, A.R.; Quadrilatero, J. Effect of acute and chronic autophagy deficiency on skeletal muscle apoptotic signaling, morphology, and function. Biochim. Biophys. Acta Mol. Cell. Res. 2017, 1864, 708–718. [Google Scholar] [CrossRef]
  16. Leduc-Gaudet, J.P.; Miguez, K.; Cefis, M.; Faitg, J.; Moamer, A.; Chaffer, T.J.; Reynaud, O.; Broering, F.E.; Shams, A.; Mayaki, D.; et al. Autophagy ablation in skeletal muscles worsens sepsis-induced muscle wasting, impairs whole-body metabolism, and decreases survival. iScience 2023, 26, 107475. [Google Scholar] [CrossRef]
  17. Doerr, V.; Montalvo, R.N.; Kwon, O.S.; Talbert, E.E.; Hain, B.A.; Houston, F.E.; Smuder, A.J. Prevention of Doxorubicin-Induced Autophagy Attenuates Oxidative Stress and Skeletal Muscle Dysfunction. Antioxidants 2020, 9, 263. [Google Scholar] [CrossRef] [PubMed]
  18. Kesavardhana, S.; Malireddi, R.K.S.; Kanneganti, T.D. Caspases in Cell Death, Inflammation, and Pyroptosis. Annu. Rev. Immunol. 2020, 38, 567–595. [Google Scholar] [CrossRef] [PubMed]
  19. Wang, B.; Zhang, C.; Zhang, A.; Cai, H.; Price, S.R.; Wang, X.H. MicroRNA-23a and MicroRNA-27a Mimic Exercise by Ameliorating CKD-Induced Muscle Atrophy. J. Am. Soc. Nephrol. 2017, 28, 2631–2640. [Google Scholar] [CrossRef]
  20. Zhang, W.; Sun, W.; Gu, X.; Miao, C.; Feng, L.; Shen, Q.; Liu, X.; Zhang, X. GDF-15 in tumor-derived exosomes promotes muscle atrophy via Bcl-2/caspase-3 pathway. Cell Death Discov. 2022, 8, 162. [Google Scholar]
  21. Zhu, S.; Nagashima, M.; Khan, M.A.; Yasuhara, S.; Kaneki, M.; Martyn, J.A. Lack of caspase-3 attenuates immobilization-induced muscle atrophy and loss of tension generation along with mitigation of apoptosis and inflammation. Muscle Nerve 2013, 47, 711–721. [Google Scholar] [CrossRef] [PubMed]
  22. Plant, P.J.; Bain, J.R.; Correa, J.E.; Woo, M.; Batt, J. Absence of caspase-3 protects against denervation-induced skeletal muscle atrophy. J. Appl. Physiol. 2009, 107, 224–234. [Google Scholar] [CrossRef] [PubMed]
  23. Van Ba, H.; Hwang, I. Role of caspase-9 in the effector caspases and genome expressions, and growth of bovine skeletal myoblasts. Dev. Growth Differ. 2014, 56, 131–142. [Google Scholar] [CrossRef] [PubMed]
  24. Whitmore, C.; Pratt, E.P.S.; Anderson, L.; Bradley, K.; Latour, S.M.; Hashmi, M.N.; Urazaev, A.K.; Weilbaecher, R.; Davie, J.K.; Wang, W.H.; et al. The ERG1a potassium channel increases basal intracellular calcium concentration and calpain activity in skeletal muscle cells. Skelet. Muscle 2020, 10, 1. [Google Scholar] [CrossRef] [PubMed]
  25. McClung, J.M.; Judge, A.R.; Talbert, E.E.; Powers, S.K. Calpain-1 is required for hydrogen peroxide-induced myotube atrophy. Am. J. Physiol. Cell Physiol. 2009, 296, C363–C371. [Google Scholar] [CrossRef] [PubMed]
  26. Talbert, E.E.; Smuder, A.J.; Min, K.; Kwon, O.S.; Powers, S.K. Calpain and caspase-3 play required roles in immobilization-induced limb muscle atrophy. J. Appl. Physiol. 2013, 114, 1482–1489. [Google Scholar] [CrossRef] [PubMed]
  27. Sharma, B.; Dutt, V.; Kaur, N.; Mittal, A.; Dabur, R. Tinospora cordifolia protects from skeletal muscle atrophy by alleviating oxidative stress and inflammation induced by sciatic denervation. J. Ethnopharmacol. 2020, 254, 112720. [Google Scholar] [CrossRef] [PubMed]
  28. Yadav, A.; Singh, A.; Phogat, J.; Dahuja, A.; Dabur, R. Magnoflorine prevent the skeletal muscle atrophy via Akt/mTOR/FoxO signal pathway and increase slow-MyHC production in streptozotocin-induced diabetic rats. J. Ethnopharmacol. 2021, 267, 113510. [Google Scholar] [CrossRef] [PubMed]
  29. Dutt, V.; Saini, V.; Gupta, P.; Kaur, N.; Bala, M.; Gujar, R.; Grewal, A.; Gupta, S.; Dua, A.; Mittal, A. S-allyl cysteine inhibits TNFalpha-induced skeletal muscle wasting through suppressing proteolysis and expression of inflammatory molecules. Biochim. Biophys. Acta Gen. Subj. 2018, 1862, 895–906. [Google Scholar] [CrossRef]
  30. Yoshida, T.; Delafontaine, P. Mechanisms of IGF-1-Mediated Regulation of Skeletal Muscle Hypertrophy and Atrophy. Cells 2020, 9, 1970. [Google Scholar] [CrossRef]
  31. Schmitt, T.L.; Martignoni, M.E.; Bachmann, J.; Fechtner, K.; Friess, H.; Kinscherf, R.; Hildebrandt, W. Activity of the Akt-dependent anabolic and catabolic pathways in muscle and liver samples in cancer-related cachexia. J. Mol. Med. 2007, 85, 647–654. [Google Scholar] [CrossRef]
  32. Leger, B.; Derave, W.; De Bock, K.; Hespel, P.; Russell, A.P. Human sarcopenia reveals an increase in SOCS-3 and myostatin and a reduced efficiency of Akt phosphorylation. Rejuvenation Res. 2008, 11, 163B–175B. [Google Scholar] [CrossRef]
  33. Goncalves, M.D.; Pistilli, E.E.; Balduzzi, A.; Birnbaum, M.J.; Lachey, J.; Khurana, T.S.; Ahima, R.S. Akt deficiency attenuates muscle size and function but not the response to ActRIIB inhibition. PLoS ONE 2010, 5, e12707. [Google Scholar] [CrossRef]
  34. Chen, M.; Ji, C.; Yang, Q.; Gao, S.; Peng, Y.; Li, Z.; Gao, X.; Li, Y.; Jiang, N.; Zhang, Y.; et al. AKT2 regulates development and metabolic homeostasis via AMPK-depedent pathway in skeletal muscle. Clin. Sci. 2020, 134, 2381–2398. [Google Scholar] [CrossRef]
  35. Tuttle, C.S.L.; Thang, L.A.N.; Maier, A.B. Markers of inflammation and their association with muscle strength and mass: A systematic review and meta-analysis. Ageing Res. Rev. 2020, 64, 101185. [Google Scholar] [CrossRef]
  36. Schaap, L.A.; Pluijm, S.M.; Deeg, D.J.; Harris, T.B.; Kritchevsky, S.B.; Newman, A.B.; Colbert, L.H.; Pahor, M.; Rubin, S.M.; Tylavsky, F.A.; et al. Higher inflammatory marker levels in older persons: Associations with 5-year change in muscle mass and muscle strength. J. Gerontol. A Biol. Sci. Med. Sci. 2009, 64, 1183–1189. [Google Scholar] [CrossRef]
  37. Tao, W.; Ouyang, Z.; Liao, Z.; Li, L.; Zhang, Y.; Gao, J.; Ma, L.; Yu, S. Ursolic Acid Alleviates Cancer Cachexia and Prevents Muscle Wasting via Activating SIRT1. Cancers 2023, 15, 2378. [Google Scholar] [CrossRef]
  38. Fang, W.Y.; Tseng, Y.T.; Lee, T.Y.; Fu, Y.C.; Chang, W.H.; Lo, W.W.; Lin, C.L.; Lo, Y.C. Triptolide prevents LPS-induced skeletal muscle atrophy via inhibiting NF-kappaB/TNF-alpha and regulating protein synthesis/degradation pathway. Br. J. Pharmacol. 2021, 178, 2998–3016. [Google Scholar] [CrossRef]
  39. Baek, J.S.; Shin, Y.J.; Ma, X.; Park, H.S.; Hwang, Y.H.; Kim, D.H. Bifidobacterium bifidum and Lactobacillus paracasei alleviate sarcopenia and cognitive impairment in aged mice by regulating gut microbiota-mediated AKT, NF-kappaB, and FOXO3a signaling pathways. Immun. Ageing 2023, 20, 56. [Google Scholar] [CrossRef]
  40. 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]
  41. Rhoads, M.G.; Kandarian, S.C.; Pacelli, F.; Doglietto, G.B.; Bossola, M. Expression of NF-kappaB and IkappaB proteins in skeletal muscle of gastric cancer patients. Eur. J. Cancer 2010, 46, 191–197. [Google Scholar] [CrossRef]
  42. Hunter, R.B.; Kandarian, S.C. Disruption of either the Nfkb1 or the Bcl3 gene inhibits skeletal muscle atrophy. J. Clin. Investig. 2004, 114, 1504–1511. [Google Scholar] [CrossRef]
  43. Thomson, D.M. The Role of AMPK in the Regulation of Skeletal Muscle Size, Hypertrophy, and Regeneration. Int. J. Mol. Sci. 2018, 19, 3125. [Google Scholar] [CrossRef]
  44. Livshits, G.; Kalinkovich, A. Restoration of epigenetic impairment in the skeletal muscle and chronic inflammation resolution as a therapeutic approach in sarcopenia. Ageing Res. Rev. 2024, 96, 102267. [Google Scholar] [CrossRef]
  45. Raun, S.H.; Ali, M.S.; Han, X.; Henríquez-Olguín, C.; Pham, T.C.P.; Meneses-Valdés, R.; Knudsen, J.R.; Willemsen, A.C.H.; Larsen, S.; Jensen, T.E.; et al. Adenosine monophosphate-activated protein kinase is elevated in human cachectic muscle and prevents cancer-induced metabolic dysfunction in mice. J. Cachexia Sarcopenia Muscle 2023, 14, 1631–1647. [Google Scholar] [CrossRef]
  46. Petrocelli, J.J.; Liu, J.; Yee, E.M.; Ferrara, P.J.; Bourrant, P.E.; de Hart, N.M.M.P.; Tatum, S.M.; Holland, W.J.; Funai, K. Skeletal muscle-specific inducible AMPKalpha1/alpha2 knockout mice develop muscle weakness, glycogen depletion, and fibrosis that persists during disuse atrophy. Am. J. Physiol. Endocrinol. Metab. 2024, 326, E50–E60. [Google Scholar] [CrossRef]
  47. Guo, Y.; Meng, J.; Tang, Y.; Wang, T.; Wei, B.; Feng, R.; Gong, B.; Wang, H.; Ji, G.; Lu, Z. AMP-activated kinase alpha2 deficiency protects mice from denervation-induced skeletal muscle atrophy. Arch. Biochem. Biophys. 2016, 600, 56–60. [Google Scholar] [CrossRef]
  48. Lee, S.J. Targeting the myostatin signaling pathway to treat muscle loss and metabolic dysfunction. J. Clin. Investig. 2021, 131, e148372. [Google Scholar] [CrossRef]
  49. Mitra, A.; Qaisar, R.; Bose, B.; Sudheer, S.P. The elusive role of myostatin signaling for muscle regeneration and maintenance of muscle and bone homeostasis. Osteoporos Sarcopenia 2023, 9, 1–7. [Google Scholar] [CrossRef]
  50. Aversa, Z.; Bonetto, A.; Penna, F.; Costelli, P.; Di Rienzo, G.; Lacitignola, A.; Baccino, F.M.; Ziparo, V.; Mercantini, P.; Rossi Fanelli, F.; et al. Changes in myostatin signaling in non-weight-losing cancer patients. Ann. Surg. Oncol. 2012, 19, 1350–1356. [Google Scholar] [CrossRef]
  51. Kawaguchi, Y.; Watanabe, A.; Shiratori, T.; Kaku, R.; Ueda, K.; Okamoto, K.; Kataoka, Y.; Ohshio, Y.; Hanaoka, J. Myostatin expression in lung cancer induces sarcopenia and promotes cancer progression. Gen. Thorac. Cardiovasc. Surg. 2024, 72, 232–239. [Google Scholar] [CrossRef]
  52. Choi, K.; Jang, H.Y.; Ahn, J.M.; Hwang, S.H.; Chung, J.W.; Choi, Y.S.; Kim, J.W.; Jang, E.S.; Choi, G.H.; Jeong, S.H. The association of the serum levels of myostatin, follistatin, and interleukin-6 with sarcopenia, and their impacts on survival in patients with hepatocellular carcinoma. Clin. Mol. Hepatol. 2020, 26, 492–505. [Google Scholar] [CrossRef]
  53. Nishikawa, H.; Enomoto, H.; Ishii, A.; Iwata, Y.; Miyamoto, Y.; Ishii, N.; Yuri, Y.; Hasegawa, K.; Nakano, C.; Nishimura, T.; et al. Elevated serum myostatin level is associated with worse survival in patients with liver cirrhosis. J. Cachexia Sarcopenia Muscle 2017, 8, 915–925. [Google Scholar] [CrossRef]
  54. Ryan, A.S.; Li, G. Skeletal muscle myostatin gene expression and sarcopenia in overweight and obese middle-aged and older adults. JCSM Clin. Rep. 2021, 6, 137–142. [Google Scholar] [CrossRef]
  55. Chew, J.; Tay, L.; Lim, J.P.; Leung, B.P.; Yeo, A.; Yew, S.; Ding, Y.Y.; Lim, W.S. Serum Myostatin and IGF-1 as Gender-Specific Biomarkers of Frailty and Low Muscle Mass in Community-Dwelling Older Adults. J. Nutr. Health Aging 2019, 23, 979–986. [Google Scholar] [CrossRef]
  56. Godala, M.; Gaszynska, E.; Walczak, K.; Malecka-Wojciesko, E. Myostatin and Activin A as Biomarkers of Sarcopenia in Inflammatory Bowel Disease Patients. Nutrients 2024, 16, 810. [Google Scholar] [CrossRef]
  57. Gilson, H.; Schakman, O.; Combaret, L.; Lause, P.; Grobet, L.; Attaix, D.; Ketelslegers, J.M.; Thissen, J.P. Myostatin gene deletion prevents glucocorticoid-induced muscle atrophy. Endocrinology 2007, 148, 452–460. [Google Scholar] [CrossRef]
  58. Mendias, C.L.; Bakhurin, K.I.; Gumucio, J.P.; Shallal-Ayzin, M.V.; Davis, C.S.; Faulkner, J.A. Haploinsufficiency of myostatin protects against aging-related declines in muscle function and enhances the longevity of mice. Aging Cell 2015, 14, 704–706. [Google Scholar] [CrossRef]
  59. Morissette, M.R.; Cook, S.A.; Buranasombati, C.; Rosenberg, M.A.; Rosenzweig, A. Myostatin inhibits IGF-I-induced myotube hypertrophy through Akt. Am. J. Physiol. Cell Physiol. 2009, 297, C1124–C1132. [Google Scholar] [CrossRef]
  60. Giannesini, B.; Vilmen, C.; Amthor, H.; Bernard, M.; Bendahan, D. Lack of myostatin impairs mechanical performance and ATP cost of contraction in exercising mouse gastrocnemius muscle In Vivo. Am. J. Physiol. Endocrinol. Metab. 2013, 305, E33–E40. [Google Scholar] [CrossRef]
  61. Dai, Z.; Ramesh, V.; Locasale, J.W. The evolving metabolic landscape of chromatin biology and epigenetics. Nat. Rev. Genet. 2020, 21, 737–753. [Google Scholar] [CrossRef]
  62. Nevi, L.; Pollanen, N.; Penna, F.; Caretti, G. Targeting Epigenetic Regulators with HDAC and BET Inhibitors to Modulate Muscle Wasting. Int. J. Mol. Sci. 2023, 24, 16404. [Google Scholar] [CrossRef]
  63. Zhao, L.Y.; Song, J.; Liu, Y.; Song, C.X.; Yi, C. Mapping the epigenetic modifications of DNA and RNA. Protein Cell 2020, 11, 792–808. [Google Scholar]
  64. Li, J.W.; Shen, Z.K.; Lin, Y.S.; Wang, Z.Y.; Li, M.L.; Sun, H.X.; Wang, Q.; Zhao, C.; Xu, J.S.; Lu, X.; et al. DNA methylation of skeletal muscle function-related secretary factors identifies FGF2 as a potential biomarker for sarcopenia. J. Cachexia Sarcopenia Muscle 2024, 15, 1209–1217. [Google Scholar]
  65. He, L.; Khanal, P.; Morse, C.I.; Williams, A.; Thomis, M. Associations of combined genetic and epigenetic scores with muscle size and muscle strength: A pilot study in older women. J. Cachexia Sarcopenia Muscle 2020, 11, 1548–1561. [Google Scholar]
  66. He, L.; Khanal, P.; Morse, C.I.; Williams, A.; Thomis, M. Differentially methylated gene patterns between age-matched sarcopenic and non-sarcopenic women. J. Cachexia Sarcopenia Muscle 2019, 10, 1295–1306. [Google Scholar]
  67. Jin, L.; Jiang, Z.; Xia, Y.; Lou, P.; Chen, L.; Wang, H.; Bai, L.; Xie, Y.; Liu, Y.; Li, W.; et al. Genome-wide DNA methylation changes in skeletal muscle between young and middle-aged pigs. BMC Genom. 2014, 15, 653. [Google Scholar]
  68. Fisher, A.G.; Seaborne, R.A.; Hughes, T.M.; Gutteridge, A.; Stewart, C.; Coulson, J.M.; Sharples, A.P.; Jarvis, J.C. Transcriptomic and epigenetic regulation of disuse atrophy and the return to activity in skeletal muscle. FASEB J. 2017, 31, 5268–5282. [Google Scholar]
  69. Tomiga, Y.; Ito, A.; Sudo, M.; Ando, S.; Eshima, H.; Sakai, K.; Nakashima, S.; Uehara, Y.; Tanaka, H.; Soejima, H.; et al. One week, but not 12 hours, of cast immobilization alters promotor DNA methylation patterns in the nNOS gene in mouse skeletal muscle. J. Physiol. 2019, 597, 5145–5159. [Google Scholar] [CrossRef]
  70. Seaborne, R.A.; Hughes, D.C.; Turner, D.C.; Owens, D.J.; Baehr, L.M.; Gorski, P.; Semenova, E.A.; Borisov, O.V.; Larin, A.K.; Popov, D.V.; et al. UBR5 is a novel E3 ubiquitin ligase involved in skeletal muscle hypertrophy and recovery from atrophy. J. Physiol. 2019, 597, 3727–3749. [Google Scholar] [CrossRef]
  71. Tajrishi, M.M.; Shin, J.; Hetman, M.; Kumar, A. DNA methyltransferase 3a and mitogen-activated protein kinase signaling regulate the expression of fibroblast growth factor-inducible 14 (Fn14) during denervation-induced skeletal muscle atrophy. J. Biol. Chem. 2014, 289, 19985–19999. [Google Scholar] [CrossRef]
  72. Hatazawa, Y.; Ono, Y.; Hirose, Y.; Kanai, S.; Fujii, N.L.; Machida, S.; Nishino, I.; Shimizu, T.; Okano, M.; Kamei, Y.; et al. Reduced Dnmt3a increases Gdf5 expression with suppressed satellite cell differentiation and impaired skeletal muscle regeneration. FASEB J. 2018, 32, 1452–1467. [Google Scholar] [CrossRef]
  73. Park, J.; Lee, K.; Kim, K.; Yi, S.J. The role of histone modifications: From neurodevelopment to neurodiseases. Signal Transduct. Target. Ther. 2022, 7, 217. [Google Scholar] [CrossRef]
  74. Xu, Y.; Zhu, Q. Histone Modifications Represent a Key Epigenetic Feature of Epithelial-to-Mesenchyme Transition in Pancreatic Cancer. Int. J. Mol. Sci. 2023, 24, 4820. [Google Scholar] [CrossRef]
  75. Liu, Y.; Wang, J.; Liu, B.; Xu, Z.Y. Dynamic regulation of DNA methylation and histone modifications in response to abiotic stresses in plants. J. Integr. Plant Biol. 2022, 64, 2252–2274. [Google Scholar] [CrossRef]
  76. Tang, J.; Zhuang, S. Histone acetylation and DNA methylation in ischemia/reperfusion injury. Clin. Sci. 2019, 133, 597–609. [Google Scholar] [CrossRef]
  77. Chelladurai, P.; Boucherat, O.; Stenmark, K.; Kracht, M.; Seeger, W.; Bauer, U.M.; Bonnet, S.; Pullamsetti, S.S. Targeting histone acetylation in pulmonary hypertension and right ventricular hypertrophy. Br. J. Pharmacol. 2021, 178, 54–71. [Google Scholar] [CrossRef]
  78. Burns, A.M.; Graff, J. Cognitive epigenetic priming: Leveraging histone acetylation for memory amelioration. Curr. Opin. Neurobiol. 2021, 67, 75–84. [Google Scholar] [CrossRef]
  79. Ohsawa, I.; Konno, R.; Masuzawa, R.; Kawano, F. Amount of daily exercise is an essential stimulation to alter the epigenome of skeletal muscle in rats. J. Appl. Physiol. 2018, 125, 1097–1104. [Google Scholar] [CrossRef]
  80. Ryder, D.J.; Judge, S.M.; Beharry, A.W.; Farnsworth, C.L.; Silva, J.C.; Judge, A.R. Identification of the Acetylation and Ubiquitin-Modified Proteome during the Progression of Skeletal Muscle Atrophy. PLoS ONE 2015, 10, e0136247. [Google Scholar] [CrossRef]
  81. Kawano, F.; Nimura, K.; Ishino, S.; Nakai, N.; Nakata, K.; Ohira, Y. Differences in histone modifications between slow- and fast-twitch muscle of adult rats and following overload, denervation, or valproic acid administration. J. Appl. Physiol. 2015, 119, 1042–1052. [Google Scholar] [CrossRef]
  82. Yoshihara, T.; Machida, S.; Tsuzuki, T.; Kakigi, R.; Chang, S.W.; Sugiura, T.; Naito, H. Age-related changes in histone modification in rat gastrocnemius muscle. Exp. Gerontol. 2019, 125, 110658. [Google Scholar] [CrossRef]
  83. Wang, J.; Feng, S.; Zhang, Q.; Qin, H.; Xu, C.; Fu, X.; Yan, L.; Zhao, Y.; Yao, K. Roles of Histone Acetyltransferases and Deacetylases in the Retinal Development and Diseases. Mol. Neurobiol. 2023, 60, 2330–2354. [Google Scholar] [CrossRef]
  84. Beharry, A.W.; Judge, A.R. Differential expression of HDAC and HAT genes in atrophying skeletal muscle. Muscle Nerve 2015, 52, 1098–1101. [Google Scholar] [CrossRef]
  85. Addicks, G.C.; Zhang, H.; Ryu, D.; Vasam, G.; Green, A.E.; Marshall, P.L.; Patel, S.; Kang, B.E.; Kim, D.; Katsyuba, E.; et al. GCN5 maintains muscle integrity by acetylating YY1 to promote dystrophin expression. J. Cell Biol. 2022, 221, e202104022. [Google Scholar] [CrossRef]
  86. Lee, D.; Goldberg, A.L. Muscle Wasting in Fasting Requires Activation of NF-kappaB and Inhibition of AKT/Mechanistic Target of Rapamycin (mTOR) by the Protein Acetylase, GCN5. J. Biol. Chem. 2015, 290, 30269–30279. [Google Scholar] [CrossRef]
  87. Chamberlain, W.; Gonnella, P.; Alamdari, N.; Aversa, Z.; Hasselgren, P.O. Multiple muscle wasting-related transcription factors are acetylated in dexamethasone-treated muscle cells. Biochem. Cell Biol. 2012, 90, 200–208. [Google Scholar] [CrossRef]
  88. Fan, Z.; Wu, J.; Chen, Q.N.; Lyu, A.K.; Chen, J.L.; Sun, Y.; Lyu, Q.; Zhao, Y.X.; Guo, A.; Liao, Z.Y.; et al. Type 2 diabetes-induced overactivation of P300 contributes to skeletal muscle atrophy by inhibiting autophagic flux. Life Sci. 2020, 258, 118243. [Google Scholar] [CrossRef]
  89. Alamdari, N.; Smith, I.J.; Aversa, Z.; Hasselgren, P.O. Sepsis and glucocorticoids upregulate p300 and downregulate HDAC6 expression and activity in skeletal muscle. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2010, 299, R509–R520. [Google Scholar] [CrossRef]
  90. Senf, S.M.; Sandesara, P.B.; Reed, S.A.; Judge, A.R. p300 Acetyltransferase activity differentially regulates the localization and activity of the FOXO homologues in skeletal muscle. Am. J. Physiol. Cell Physiol. 2011, 300, C1490–C1501. [Google Scholar] [CrossRef]
  91. Sin, T.K.; Zhang, G.; Zhang, Z.; Zhu, J.Z.; Zuo, Y.; Frost, J.A.; Li, M.; Li, Y.P. Cancer-Induced Muscle Wasting Requires p38beta MAPK Activation of p300. Cancer Res. 2021, 81, 885–897. [Google Scholar] [CrossRef]
  92. Sin, T.K.; Zhu, J.Z.; Zhang, G.; Li, Y.P. p300 Mediates Muscle Wasting in Lewis Lung Carcinoma. Cancer Res. 2019, 79, 1331–1342. [Google Scholar] [CrossRef]
  93. Wang, X.; Waschke, B.C.; Woolaver, R.A.; Chen, S.M.Y.; Chen, Z.; Wang, J.H. HDAC inhibitors overcome immunotherapy resistance in B-cell lymphoma. Protein Cell 2020, 11, 472–482. [Google Scholar] [CrossRef]
  94. Moresi, V.; Williams, A.H.; Meadows, E.; Flynn, J.M.; Potthoff, M.J.; McAnally, J.; Shelton, J.M.; Backs, J.; Klein, W.H.; Richardson, J.A.; et al. Myogenin and class II HDACs control neurogenic muscle atrophy by inducing E3 ubiquitin ligases. Cell 2010, 143, 35–45. [Google Scholar] [CrossRef]
  95. Luo, L.; Martin, S.C.; Parkington, J.; Cadena, S.M.; Zhu, J.; Ibebunjo, C.; Summermatter, S.; Londraville, N.; Patora-Komisarska, K.; Widler, L.; et al. HDAC4 Controls Muscle Homeostasis through Deacetylation of Myosin Heavy Chain, PGC-1alpha, and Hsc70. Cell Rep. 2019, 29, 749–763 e712. [Google Scholar]
  96. Pigna, E.; Simonazzi, E.; Sanna, K.; Bernadzki, K.M.; Proszynski, T.; Heil, C.; Palacios, D.; Adamo, S.; Moresi, V. Histone deacetylase 4 protects from denervation and skeletal muscle atrophy in a murine model of amyotrophic lateral sclerosis. EBioMedicine 2019, 40, 717–732. [Google Scholar] [CrossRef]
  97. Beharry, A.W.; Sandesara, P.B.; Roberts, B.M.; Ferreira, L.F.; Senf, S.M.; Judge, A.R. HDAC1 activates FoxO and is both sufficient and required for skeletal muscle atrophy. J. Cell Sci. 2014, 127 Pt 7, 1441–1453. [Google Scholar] [CrossRef]
  98. Ratti, F.; Ramond, F.; Moncollin, V.; Simonet, T.; Milan, G.; Mejat, A.; Thomas, J.L.; Streichenberger, N.; Gilquin, B.; Matthias, P.; et al. Histone deacetylase 6 is a FoxO transcription factor-dependent effector in skeletal muscle atrophy. J. Biol. Chem. 2015, 290, 4215–4224. [Google Scholar] [CrossRef]
  99. Ding, J.; Li, F.; Cong, Y.; Miao, J.; Wu, D.; Liu, B.; Wang, L. Trichostatin A inhibits skeletal muscle atrophy induced by cigarette smoke exposure in mice. Life Sci. 2019, 235, 116800. [Google Scholar] [CrossRef]
  100. Yoo, Y.E.; Ko, C.P. Treatment with trichostatin A initiated after disease onset delays disease progression and increases survival in a mouse model of amyotrophic lateral sclerosis. Exp. Neurol. 2011, 231, 147–159. [Google Scholar] [CrossRef]
  101. Dupre-Aucouturier, S.; Castells, J.; Freyssenet, D.; Desplanches, D. Trichostatin A, a histone deacetylase inhibitor, modulates unloaded-induced skeletal muscle atrophy. J. Appl. Physiol. 2015, 119, 342–351. [Google Scholar] [CrossRef]
  102. Walsh, M.E.; Bhattacharya, A.; Sataranatarajan, K.; Qaisar, R.; Sloane, L.; Rahman, M.M.; Kinter, M.; Van Remmen, H. The histone deacetylase inhibitor butyrate improves metabolism and reduces muscle atrophy during aging. Aging Cell 2015, 14, 957–970. [Google Scholar] [CrossRef] [PubMed]
  103. Mochalova, E.P.; Belova, S.P.; Kostrominova, T.Y.; Shenkman, B.S.; Nemirovskaya, T.L. Differences in the Role of HDACs 4 and 5 in the Modulation of Processes Regulating MAFbx and MuRF1 Expression during Muscle Unloading. Int. J. Mol. Sci. 2020, 21, 4815. [Google Scholar] [CrossRef] [PubMed]
  104. Osseni, A.; Ravel-Chapuis, A.; Belotti, E.; Scionti, I.; Gangloff, Y.G.; Moncollin, V.; Mazelin, L.; Mounier, R.; Leblanc, P.; Jasmin, B.J.; et al. Pharmacological inhibition of HDAC6 improves muscle phenotypes in dystrophin-deficient mice by downregulating TGF-beta via Smad3 acetylation. Nat. Commun. 2022, 13, 7108. [Google Scholar] [CrossRef] [PubMed]
  105. Sun, R.; Zhang, S.; Hu, W.; Lu, X.; Lou, N.; Yang, Z.; Chen, S.; Zhang, X.; Yang, H. Valproic acid attenuates skeletal muscle wasting by inhibiting C/EBPbeta-regulated atrogin1 expression in cancer cachexia. Am. J. Physiol. Cell Physiol. 2016, 311, C101–C115. [Google Scholar] [CrossRef] [PubMed]
  106. Tan, H.W.; Sim, A.Y.; Huang, S.L.; Leng, Y.; Long, Y.C. HC toxin (a HDAC inhibitor) enhances IRS1-Akt signalling and metabolism in mouse myotubes. J. Mol. Endocrinol. 2015, 55, 197–207. [Google Scholar] [CrossRef] [PubMed]
  107. Habibian, J.S.; Bolino, M.; Qian, A.; Woolsey, R.; Quilici, D.; Petereit, J.; Ferguson, B.S. Class I HDAC inhibitors attenuate dexamethasone-induced muscle atrophy via increased protein kinase C (PKC) delta phosphorylation. Cell. Signal. 2023, 110, 110815. [Google Scholar] [CrossRef] [PubMed]
  108. Bannister, A.J.; Kouzarides, T. Reversing histone methylation. Nature 2005, 436, 1103–1106. [Google Scholar] [CrossRef]
  109. Kooistra, S.M.; Helin, K. Molecular mechanisms and potential functions of histone demethylases. Nat. Rev. Mol. Cell Biol. 2012, 13, 297–311. [Google Scholar] [CrossRef] [PubMed]
  110. Bannister, A.J.; Kouzarides, T. Regulation of chromatin by histone modifications. Cell Res. 2011, 21, 381–395. [Google Scholar] [CrossRef] [PubMed]
  111. Song, Y.; Wu, F.; Wu, J. Targeting histone methylation for cancer therapy: Enzymes, inhibitors, biological activity and perspectives. J. Hematol. Oncol. 2016, 9, 49. [Google Scholar] [CrossRef]
  112. Stouth, D.W.; Manta, A.; Ljubicic, V. Protein arginine methyltransferase expression, localization, and activity during disuse-induced skeletal muscle plasticity. Am. J. Physiol. Cell Physiol. 2018, 314, C177–C190. [Google Scholar] [CrossRef] [PubMed]
  113. Fulton, M.D.; Brown, T.; Zheng, Y.G. Mechanisms and Inhibitors of Histone Arginine Methylation. Chem. Rec. 2018, 18, 1792–1807. [Google Scholar] [CrossRef] [PubMed]
  114. Choi, S.; Jeong, H.J.; Kim, H.; Choi, D.; Cho, S.C.; Seong, J.K.; Koo, S.H.; Kang, J.S. Skeletal muscle-specific Prmt1 deletion causes muscle atrophy via deregulation of the PRMT6-FOXO3 axis. Autophagy 2019, 15, 1069–1081. [Google Scholar] [CrossRef] [PubMed]
  115. Stouth, D.W.; vanLieshout, T.L.; Mikhail, A.I.; Ng, S.Y.; Raziee, R.; Edgett, B.A.; Vasam, G.; Webb, E.K.; Gilotra, K.S.; Markou, M.; et al. CARM1 drives mitophagy and autophagy flux during fasting-induced skeletal muscle atrophy. Autophagy 2024, 20, 1247–1269. [Google Scholar] [CrossRef] [PubMed]
  116. Webb, E.K.; Ng, S.Y.; Mikhail, A.I.; Stouth, D.W.; vanLieshout, T.L.; Syroid, A.L.; Ljubicic, V. Impact of short-term, pharmacological CARM1 inhibition on skeletal muscle mass, function, and atrophy in mice. Am. J. Physiol. Endocrinol. Metab. 2023, 325, E252–E266. [Google Scholar] [CrossRef] [PubMed]
  117. Kealy, L.; Runting, J.; Thiele, D.; Scheer, S. An emerging maestro of immune regulation: How DOT1L orchestrates the harmonies of the immune system. Front. Immunol. 2024, 15, 1385319. [Google Scholar] [CrossRef] [PubMed]
  118. Lakhdar, R.; Drost, E.M.; MacNee, W.; Bastos, R.; Rabinovich, R.A. 2D-DIGE proteomic analysis of vastus lateralis from COPD patients with low and normal fat free mass index and healthy controls. Respir. Res. 2017, 18, 81. [Google Scholar] [CrossRef] [PubMed]
  119. Banerjee, T.; Chakravarti, D. A peek into the complex realm of histone phosphorylation. Mol. Cell Biol. 2011, 31, 4858–4873. [Google Scholar] [CrossRef] [PubMed]
  120. Rossetto, D.; Avvakumov, N.; Cote, J. Histone phosphorylation: A chromatin modification involved in diverse nuclear events. Epigenetics 2012, 7, 1098–1108. [Google Scholar] [CrossRef]
  121. Alsharidah, M.; Lazarus, N.R.; George, T.E.; Agley, C.C.; Velloso, C.P.; Harridge, S.D. Primary human muscle precursor cells obtained from young and old donors produce similar proliferative, differentiation and senescent profiles in culture. Aging Cell 2013, 12, 333–344. [Google Scholar] [CrossRef]
  122. Huang, Y.; Wang, B.; Hassounah, F.; Price, S.R.; Klein, J.; Mohamed, T.M.A.; Wang, Y.; Park, J.; Cai, H.; Zhang, X.; et al. The impact of senescence on muscle wasting in chronic kidney disease. J. Cachexia Sarcopenia Muscle 2023, 14, 126–141. [Google Scholar] [CrossRef]
  123. Mutsaers, C.A.; Wishart, T.M.; Lamont, D.J.; Riessland, M.; Schreml, J.; Comley, L.H.; Murray, L.M.; Parson, S.H.; Lochmuller, H.; Wirth, B.; et al. Reversible molecular pathology of skeletal muscle in spinal muscular atrophy. Hum. Mol. Genet. 2011, 20, 4334–4344. [Google Scholar] [CrossRef]
  124. Gerosa, L.; Malvandi, A.M.; Gomarasca, M.; Verdelli, C.; Sansoni, V.; Faraldi, M.; Ziemann, E.; Olivieri, F.; Banfi, G.; Lombardi, G. Murine Myoblasts Exposed to SYUIQ-5 Acquire Senescence Phenotype and Differentiate into Sarcopenic-Like Myotubes, an In Vitro Study. J. Gerontol. A Biol. Sci. Med. Sci. 2024, 79, glae022. [Google Scholar] [CrossRef]
  125. Mattiroli, F.; Penengo, L. Histone Ubiquitination: An Integrative Signaling Platform in Genome Stability. Trends Genet. 2021, 37, 566–581. [Google Scholar] [CrossRef]
  126. Yu, X.; Yang, J.; Xu, J.; Pan, H.; Wang, W.; Yu, X.; Shi, S. Histone lactylation: From tumor lactate metabolism to epigenetic regulation. Int. J. Biol. Sci. 2024, 20, 1833–1854. [Google Scholar] [CrossRef]
  127. Brooks, G.A. Lactate as a fulcrum of metabolism. Redox Biol. 2020, 35, 101454. [Google Scholar] [CrossRef]
  128. Xie, Y.; Hu, H.; Liu, M.; Zhou, T.; Cheng, X.; Huang, W.; Cao, L. The role and mechanism of histone lactylation in health and diseases. Front. Genet. 2022, 13, 949252. [Google Scholar] [CrossRef]
  129. Tonkin, J.; Villarroya, F.; Puri, P.L.; Vinciguerra, M. SIRT1 signaling as potential modulator of skeletal muscle diseases. Curr. Opin. Pharmacol. 2012, 12, 372–376. [Google Scholar] [CrossRef]
  130. Lee, E.J.; Lee, M.M.; Park, S.; Jeong, K.S. Sirt2 positively regulates muscle regeneration after Notexin-induced muscle injury. Exp. Mol. Pathol. 2022, 127, 104798. [Google Scholar] [CrossRef]
  131. Zhang, M.; Tang, J.; Li, Y.; Xie, Y.; Shan, H.; Chen, M.; Zhang, J.; Yang, X.; Zhang, Q.; Yang, X. Curcumin attenuates skeletal muscle mitochondrial impairment in COPD rats: PGC-1alpha/SIRT3 pathway involved. Chem. Biol. Interact. 2017, 277, 168–175. [Google Scholar] [CrossRef]
  132. Orsolic, I.; Carrier, A.; Esteller, M. Genetic and epigenetic defects of the RNA modification machinery in cancer. Trends Genet. 2023, 39, 74–88. [Google Scholar] [CrossRef]
  133. Wang, C.; Hou, X.; Guan, Q.; Zhou, H.; Zhou, L.; Liu, L.; Liu, J.; Li, F.; Li, W.; Liu, H. RNA modification in cardiovascular disease: Implications for therapeutic interventions. Signal Transduct. Target. Ther. 2023, 8, 412. [Google Scholar] [CrossRef]
  134. An, Y.; Duan, H. The role of m6A RNA methylation in cancer metabolism. Mol. Cancer 2022, 21, 14. [Google Scholar] [CrossRef]
  135. Yang, B.; Wang, J.Q.; Tan, Y.; Yuan, R.; Chen, Z.S.; Zou, C. RNA methylation and cancer treatment. Pharmacol. Res. 2021, 174, 105937. [Google Scholar] [CrossRef]
  136. Zhang, X.; Yao, Y.; Han, J.; Yang, Y.; Chen, Y.; Tang, Z.; Gao, F. Longitudinal epitranscriptome profiling reveals the crucial role of N(6)-methyladenosine methylation in porcine prenatal skeletal muscle development. J. Genet. Genom. 2020, 47, 466–476. [Google Scholar] [CrossRef]
  137. Gheller, B.J.; Blum, J.E.; Fong, E.H.H.; Malysheva, O.V.; Cosgrove, B.D.; Thalacker-Mercer, A.E. A defined N6-methyladenosine (m(6)A) profile conferred by METTL3 regulates muscle stem cell/myoblast state transitions. Cell Death Discov. 2020, 6, 95. [Google Scholar] [CrossRef]
  138. Sun, J.; Zhou, H.; Chen, Z.; Zhang, H.; Cao, Y.; Yao, X.; Chen, X.; Liu, B.; Gao, Z.; Shen, Y.; et al. Altered m6A RNA methylation governs denervation-induced muscle atrophy by regulating ubiquitin proteasome pathway. J. Transl. Med. 2023, 21, 845. [Google Scholar] [CrossRef]
  139. Liu, Y.; Zhou, T.; Wang, Q.; Fu, R.; Zhang, Z.; Chen, N.; Li, Z.; Gao, G.; Peng, S.; Yang, D. m(6) A demethylase ALKBH5 drives denervation-induced muscle atrophy by targeting HDAC4 to activate FoxO3 signalling. J. Cachexia Sarcopenia Muscle 2022, 13, 1210–1223. [Google Scholar] [CrossRef]
  140. Kung, M.L.; Yang, T.H.; Lin, C.C.; Ho, J.Y.; Hung, T.C.; Chang, C.H.; Huang, K.W.; Chen, C.C.; Chen, Y.W. ADAR2 deficiency ameliorates non-alcoholic fatty liver disease and muscle atrophy through modulating serum amyloid A1. J. Cachexia Sarcopenia Muscle 2024, 15, 949–962. [Google Scholar] [CrossRef]
  141. Deng, L.; Han, X.; Wang, Z.; Nie, X.; Bian, J. The Landscape of Noncoding RNA in Pulmonary Hypertension. Biomolecules 2022, 12, 796. [Google Scholar] [CrossRef]
  142. van der Werf, J.; Chin, C.V.; Fleming, N.I. SnoRNA in Cancer Progression, Metastasis and Immunotherapy Response. Biology 2021, 10, 809. [Google Scholar] [CrossRef]
  143. Gagan, J.; Dey, B.K.; Layer, R.; Yan, Z.; Dutta, A. MicroRNA-378 targets the myogenic repressor MyoR during myoblast differentiation. J. Biol. Chem. 2011, 286, 19431–19438. [Google Scholar] [CrossRef]
  144. Iwasaki, H.; Ichihara, Y.; Morino, K.; Lemecha, M.; Sugawara, L.; Sawano, T.; Miake, J.; Sakurai, H.; Nishi, E.; Maegawa, H.; et al. MicroRNA-494-3p inhibits formation of fast oxidative muscle fibres by targeting E1A-binding protein p300 in human-induced pluripotent stem cells. Sci. Rep. 2021, 11, 1161. [Google Scholar] [CrossRef]
  145. Yang, X.; Xue, P.; Chen, H.; Yuan, M.; Kang, Y.; Duscher, D.; Machens, H.G.; Chen, Z. Denervation drives skeletal muscle atrophy and induces mitochondrial dysfunction, mitophagy and apoptosis via miR-142a-5p/MFN1 axis. Theranostics 2020, 10, 1415–1432. [Google Scholar] [CrossRef]
  146. Liu, C.; Li, L.; Ge, M.; Gu, L.; Zhang, K.; Su, Y.; Zhang, Y.; Liu, C.; Lan, M.; Yu, Y.; et al. MiR-29ab1 Cluster Resists Muscle Atrophy Through Inhibiting MuRF1. DNA Cell Biol. 2021, 40, 1167–1176. [Google Scholar] [CrossRef]
  147. Liang, R.; Shen, X.; Wang, F.; Wang, X.; DesJarlais, A.; Syed, A.; Saba, R.; Tan, Z.; Yu, F.; Ji, X.; et al. H19X-encoded miR-322(424)/miR-503 regulates muscle mass by targeting translation initiation factors. J. Cachexia Sarcopenia Muscle 2021, 12, 2174–2186. [Google Scholar] [CrossRef]
  148. Shao, X.; Gong, W.; Wang, Q.; Wang, P.; Shi, T.; Mahmut, A.; Qin, J.; Yao, Y.; Yan, W.; Chen, D.; et al. Atrophic skeletal muscle fibre-derived small extracellular vesicle miR-690 inhibits satellite cell differentiation during ageing. J. Cachexia Sarcopenia Muscle 2022, 13, 3163–3180. [Google Scholar] [CrossRef]
  149. Goljanek-Whysall, K.; Soriano-Arroquia, A.; McCormick, R.; Chinda, C.; McDonagh, B. miR-181a regulates p62/SQSTM1, parkin, and protein DJ-1 promoting mitochondrial dynamics in skeletal muscle aging. Aging Cell 2020, 19, e13140. [Google Scholar] [CrossRef]
  150. Soriano-Arroquia, A.; House, L.; Tregilgas, L.; Canty-Laird, E.; Goljanek-Whysall, K. The functional consequences of age-related changes in microRNA expression in skeletal muscle. Biogerontology 2016, 17, 641–654. [Google Scholar] [CrossRef]
  151. Patricia, S.P.; Ameena, H.; Aladin, M.B.; Mohamed, J.S. MicroRNA-434-3p regulates age-related apoptosis through eIF5A1 in the skeletal muscle. Aging 2017, 9, 1012–1029. [Google Scholar]
  152. Hu, Z.; Klein, J.D.; Mitch, W.E.; Zhang, L.; Martinez, I.; Wang, X.H. MicroRNA-29 induces cellular senescence in aging muscle through multiple signaling pathways. Aging 2014, 6, 160–175. [Google Scholar] [CrossRef]
  153. Chen, Z.L.; Guo, C.; Zou, Y.Y.; Feng, C.; Yang, D.X.; Sun, C.C.; Wen, W.; Jian, Z.J.; Zhao, Z.; Xiao, Q.; et al. Aerobic exercise enhances mitochondrial homeostasis to counteract D-galactose-induced sarcopenia in zebrafish. Exp. Gerontol. 2023, 180, 112265. [Google Scholar] [CrossRef]
  154. Chen, F.X.; Shen, Y.; Liu, Y.; Wang, H.F.; Liang, C.Y.; Luo, M. Inflammation-dependent downregulation of miR-532-3p mediates apoptotic signaling in human sarcopenia through targeting BAK1. Int. J. Biol. Sci. 2020, 16, 1481–1494. [Google Scholar] [CrossRef]
  155. Li, Z.; Liu, C.; Li, S.; Li, T.; Li, Y.; Wang, N.; Bao, X.; Xue, P.; Liu, S. BMSC-Derived Exosomes Inhibit Dexamethasone-Induced Muscle Atrophy via the miR-486-5p/FoxO1 Axis. Front. Endocrinol. 2021, 12, 681267. [Google Scholar] [CrossRef]
  156. Hudson, M.B.; Rahnert, J.A.; Zheng, B.; Woodworth-Hobbs, M.E.; Franch, H.A.; Price, S.R. miR-182 attenuates atrophy-related gene expression by targeting FoxO3 in skeletal muscle. Am. J. Physiol. Cell Physiol. 2014, 307, C314–C319. [Google Scholar] [CrossRef]
  157. van de Worp, W.; Schols, A.; Dingemans, A.C.; Op den Kamp, C.M.H.; Degens, J.; Kelders, M.; Coort, S.; Woodruff, H.C.; Kratassiouk, G.; Harel-Bellan, A.; et al. Identification of microRNAs in skeletal muscle associated with lung cancer cachexia. J. Cachexia Sarcopenia Muscle 2020, 11, 452–463. [Google Scholar] [CrossRef]
  158. Chang, S.Y.; Han, S.Z.; Choe, H.M.; Gao, K.; Jin, Z.Y.; Liu, X.Y.; Yang, L.H.; Lv, S.T.; Yin, X.J.; Quan, L.H.; et al. miR-320 regulates myogenesis by targeting growth factor receptor-bound protein-2 and ameliorates myotubes atrophy. Int. J. Biochem. Cell Biol. 2022, 147, 106212. [Google Scholar] [CrossRef]
  159. Li, J.; Chan, M.C.; Yu, Y.; Bei, Y.; Chen, P.; Zhou, Q.; Cheng, L.; Chen, L.; Ziegler, O.; Rowe, G.C.; et al. miR-29b contributes to multiple types of muscle atrophy. Nat. Commun. 2017, 8, 15201. [Google Scholar] [CrossRef]
  160. Wada, S.; Kato, Y.; Okutsu, M.; Miyaki, S.; Suzuki, K.; Yan, Z.; Schiaffino, S.; Asahara, H.; Ushida, T.; Akimoto, T. Translational suppression of atrophic regulators by microRNA-23a integrates resistance to skeletal muscle atrophy. J. Biol. Chem. 2011, 286, 38456–38465. [Google Scholar] [CrossRef]
  161. Yang, X.; Li, Z.; Wang, Z.; Yu, J.; Ma, M.; Nie, Q. miR-27b-3p Attenuates Muscle Atrophy by Targeting Cbl-b in Skeletal Muscles. Biomolecules 2022, 12, 191. [Google Scholar] [CrossRef]
  162. Xie, K.; Xiong, H.; Xiao, W.; Xiong, Z.; Hu, W.; Ye, J.; Xu, N.; Shi, J.; Yuan, C.; Chen, Z.; et al. Downregulation of miR-29c promotes muscle wasting by modulating the activity of leukemia inhibitory factor in lung cancer cachexia. Cancer Cell Int. 2021, 21, 627. [Google Scholar] [CrossRef]
  163. Qiu, L.; Chen, W.; Wu, C.; Yuan, Y.; Li, Y. Exosomes of oral squamous cell carcinoma cells containing miR-181a-3p induce muscle cell atrophy and apoptosis by transmissible endoplasmic reticulum stress signaling. Biochem. Biophys. Res. Commun. 2020, 533, 831–837. [Google Scholar] [CrossRef]
  164. Liu, C.; Wang, M.; Chen, M.; Zhang, K.; Gu, L.; Li, Q.; Yu, Z.; Li, N.; Meng, Q. miR-18a induces myotubes atrophy by down-regulating IgfI. Int. J. Biochem. Cell Biol. 2017, 90, 145–154. [Google Scholar] [CrossRef]
  165. Cho, K.A.; Choi, D.W.; Kim, Y.H.; Kim, J.; Ryu, K.H.; Woo, S.Y. Mesenchymal Stem Cell-Derived Exosomes Protect Muscle Loss by miR-145-5p Activity Targeting Activin A Receptors. Cells 2021, 10, 2169. [Google Scholar] [CrossRef]
  166. Zhang, A.; Wang, H.; Wang, B.; Yuan, Y.; Klein, J.D.; Wang, X.H. Exogenous miR-26a suppresses muscle wasting and renal fibrosis in obstructive kidney disease. FASEB J. 2019, 33, 13590–13601. [Google Scholar] [CrossRef]
  167. Wang, X.H.; Hu, Z.; Klein, J.D.; Zhang, L.; Fang, F.; Mitch, W.E. Decreased miR-29 suppresses myogenesis in CKD. J. Am. Soc. Nephrol. 2011, 22, 2068–2076. [Google Scholar] [CrossRef]
  168. Wang, B.; Wang, J.; He, W.; Zhao, Y.; Zhang, A.; Liu, Y.; Hassounah, F.; Ma, F.; Klein, J.D.; Wang, X.H.; et al. Exogenous miR-29a Attenuates Muscle Atrophy and Kidney Fibrosis in Unilateral Ureteral Obstruction Mice. Hum. Gene Ther. 2020, 31, 367–375. [Google Scholar] [CrossRef]
  169. Wang, H.; Wang, B.; Zhang, A.; Hassounah, F.; Seow, Y.; Wood, M.; Ma, F.; Klein, J.D.; Price, S.R.; Wang, X.H. Exosome-Mediated miR-29 Transfer Reduces Muscle Atrophy and Kidney Fibrosis in Mice. Mol. Ther. 2019, 27, 571–583. [Google Scholar] [CrossRef]
  170. McFarlane, C.; Vajjala, A.; Arigela, H.; Lokireddy, S.; Ge, X.; Bonala, S.; Manickam, R.; Kambadur, R.; Sharma, M. Negative auto-regulation of myostatin expression is mediated by Smad3 and microRNA-27. PLoS ONE 2014, 9, e87687. [Google Scholar] [CrossRef]
  171. Xu, J.; Li, R.; Workeneh, B.; Dong, Y.; Wang, X.; Hu, Z. Transcription factor FoxO1, the dominant mediator of muscle wasting in chronic kidney disease, is inhibited by microRNA-486. Kidney Int. 2012, 82, 401–411. [Google Scholar] [CrossRef]
  172. Hudson, M.B.; Woodworth-Hobbs, M.E.; Zheng, B.; Rahnert, J.A.; Blount, M.A.; Gooch, J.L.; Searles, C.D.; Price, S.R. miR-23a is decreased during muscle atrophy by a mechanism that includes calcineurin signaling and exosome-mediated export. Am. J. Physiol. Cell Physiol. 2014, 306, C551–C558. [Google Scholar] [CrossRef]
  173. Zhang, A.; Li, M.; Wang, B.; Klein, J.D.; Price, S.R.; Wang, X.H. miRNA-23a/27a attenuates muscle atrophy and renal fibrosis through muscle-kidney crosstalk. J. Cachexia Sarcopenia Muscle 2018, 9, 755–770. [Google Scholar] [CrossRef]
  174. Yang, S.; Yang, G.; Wu, H.; Kang, L.; Xiang, J.; Zheng, P.; Qiu, S.; Liang, Z.; Lu, Y.; Jia, L. MicroRNA-193b impairs muscle growth in mouse models of type 2 diabetes by targeting the PDK1/Akt signalling pathway. Diabetologia 2022, 65, 563–581. [Google Scholar] [CrossRef]
  175. Ikenaka, A.; Kitagawa, Y.; Yoshida, M.; Lin, C.Y.; Niwa, A.; Nakahata, T.; Saito, M.K. SMN promotes mitochondrial metabolic maturation during myogenesis by regulating the MYOD-miRNA axis. Life Sci. Alliance 2023, 6, e202201457. [Google Scholar] [CrossRef]
  176. Pourshafie, N.; Lee, P.R.; Chen, K.L.; Harmison, G.G.; Bott, L.C.; Fischbeck, K.H.; Rinaldi, C. Systemic Delivery of MicroRNA Using Recombinant Adeno-associated Virus Serotype 9 to Treat Neuromuscular Diseases in Rodents. J. Vis. Exp. 2018, 138, e55724. [Google Scholar]
  177. Boon, H.; Sjogren, R.J.; Massart, J.; Egan, B.; Kostovski, E.; Iversen, P.O.; Hjeltnes, N.; Chibalin, A.V.; Widegren, U.; Zierath, J.R. MicroRNA-208b progressively declines after spinal cord injury in humans and is inversely related to myostatin expression. Physiol. Rep. 2015, 3, e12622. [Google Scholar] [CrossRef]
  178. Van Pelt, D.W.; Vechetti, I.J., Jr.; Lawrence, M.M.; Van Pelt, K.L.; Patel, P.; Miller, B.F.; Butterfield, T.A.; Dupont-Versteegden, E.E. Serum extracellular vesicle miR-203a-3p content is associated with skeletal muscle mass and protein turnover during disuse atrophy and regrowth. Am. J. Physiol. Cell Physiol. 2020, 319, C419–C431. [Google Scholar] [CrossRef]
  179. McCarthy, J.J.; Esser, K.A.; Peterson, C.A.; Dupont-Versteegden, E.E. Evidence of MyomiR network regulation of beta-myosin heavy chain gene expression during skeletal muscle atrophy. Physiol. Genom. 2009, 39, 219–226. [Google Scholar] [CrossRef]
  180. Jung, W.; Juang, U.; Gwon, S.; Nguyen, H.; Huang, Q.; Lee, S.; Lee, B.; Kim, S.H.; Ryu, S.; Park, J.; et al. Identifying the potential therapeutic effects of miR-6516 on muscle disuse atrophy. Mol. Med. Rep. 2024, 30, 119. [Google Scholar] [CrossRef]
  181. Khayrullin, A.; Smith, L.; Mistry, D.; Dukes, A.; Pan, Y.A.; Hamrick, M.W. Chronic alcohol exposure induces muscle atrophy (myopathy) in zebrafish and alters the expression of microRNAs targeting the Notch pathway in skeletal muscle. Biochem. Biophys. Res. Commun. 2016, 479, 590–595. [Google Scholar] [CrossRef]
  182. Shin, J.; Miyaki, S.; Asahara, H.; Akimoto, T. MicroRNA-140 is not involved in sepsis-induced muscle atrophy. Am. J. Physiol. Cell Physiol. 2023, 325, C509–C518. [Google Scholar] [CrossRef]
  183. Borja-Gonzalez, M.; Casas-Martinez, J.C.; McDonagh, B.; Goljanek-Whysall, K. Inflamma-miR-21 Negatively Regulates Myogenesis during Ageing. Antioxidants 2020, 9, 345. [Google Scholar] [CrossRef]
  184. Freire, P.P.; Cury, S.S.; Lopes, L.O.; Fernandez, G.J.; Liu, J.; de Moraes, L.N.; de Oliveira, G.; Oliveira, J.S.; de Moraes, D.; Cabral-Marques, O.; et al. Decreased miR-497-5p Suppresses IL-6 Induced Atrophy in Muscle Cells. Cells 2021, 10, 3527. [Google Scholar] [CrossRef]
  185. Che, J.; Xu, C.; Wu, Y.; Jia, P.; Han, Q.; Ma, Y.; Wang, X.; Zheng, Y. MiR-1290 promotes myoblast differentiation and protects against myotube atrophy via Akt/p70/FoxO3 pathway regulation. Skelet Muscle 2021, 11, 6. [Google Scholar] [CrossRef]
  186. Puig-Vilanova, E.; Ausin, P.; Martinez-Llorens, J.; Gea, J.; Barreiro, E. Do epigenetic events take place in the vastus lateralis of patients with mild chronic obstructive pulmonary disease? PLoS ONE 2014, 9, e102296. [Google Scholar]
  187. Farre-Garros, R.; Lee, J.Y.; Natanek, S.A.; Connolly, M.; Sayer, A.A.; Patel, H.; Cooper, C.; Polkey, M.I.; Kemp, P.R. Quadriceps miR-542-3p and -5p are elevated in COPD and reduce function by inhibiting ribosomal and protein synthesis. J. Appl. Physiol. 2019, 126, 1514–1524. [Google Scholar] [CrossRef]
  188. Garros, R.F.; Paul, R.; Connolly, M.; Lewis, A.; Garfield, B.E.; Natanek, S.A.; Bloch, S.; Mouly, V.; Griffiths, M.J.; Polkey, M.I.; et al. MicroRNA-542 Promotes Mitochondrial Dysfunction and SMAD Activity and Is Elevated in Intensive Care Unit-acquired Weakness. Am. J. Respir. Crit. Care Med. 2017, 196, 1422–1433. [Google Scholar] [CrossRef]
  189. Connolly, M.; Paul, R.; Farre-Garros, R.; Natanek, S.A.; Bloch, S.; Lee, J.; Lorenzo, J.P.; Patel, H.; Cooper, C.; Sayer, A.A.; et al. miR-424-5p reduces ribosomal RNA and protein synthesis in muscle wasting. J. Cachexia Sarcopenia Muscle 2018, 9, 400–416. [Google Scholar] [CrossRef]
  190. Jin, J.; Li, F.; Fan, C.; Wu, Y.; He, C. Elevated mir-145-5p is associated with skeletal muscle dysfunction and triggers apoptotic cell death in C2C12 myotubes. J. Muscle Res. Cell Motil. 2022, 43, 135–145. [Google Scholar] [CrossRef]
  191. Lewis, A.; Riddoch-Contreras, J.; Natanek, S.A.; Donaldson, A.; Man, W.D.; Moxham, J.; Hopkinson, N.S.; Polkey, M.I.; Kemp, P.R. Downregulation of the serum response factor/miR-1 axis in the quadriceps of patients with COPD. Thorax 2012, 67, 26–34. [Google Scholar] [CrossRef]
  192. Williams, A.H.; Valdez, G.; Moresi, V.; Qi, X.; McAnally, J.; Elliott, J.L.; Bassel-Duby, R.; Sanes, J.R.; Olson, E.N. MicroRNA-206 delays ALS progression and promotes regeneration of neuromuscular synapses in mice. Science 2009, 326, 1549–1554. [Google Scholar] [CrossRef]
  193. Guess, M.G.; Barthel, K.K.; Harrison, B.C.; Leinwand, L.A. miR-30 family microRNAs regulate myogenic differentiation and provide negative feedback on the microRNA pathway. PLoS ONE 2015, 10, e0118229. [Google Scholar] [CrossRef]
  194. Klockner, I.; Schutt, C.; Gerhardt, T.; Boettger, T.; Braun, T. Control of CRK-RAC1 activity by the miR-1/206/133 miRNA family is essential for neuromuscular junction function. Nat. Commun. 2022, 13, 3180. [Google Scholar] [CrossRef]
  195. Schutt, C.; Hallmann, A.; Hachim, S.; Klockner, I.; Valussi, M.; Atzberger, A.; Graumann, J.; Braun, T.; Boettger, T. Linc-MYH configures INO80 to regulate muscle stem cell numbers and skeletal muscle hypertrophy. EMBO J. 2020, 39, e105098. [Google Scholar] [CrossRef]
  196. Lv, W.; Jiang, W.; Luo, H.; Tong, Q.; Niu, X.; Liu, X.; Miao, Y.; Wang, J.; Guo, Y.; Li, J.; et al. Long noncoding RNA lncMREF promotes myogenic differentiation and muscle regeneration by interacting with the Smarca5/p300 complex. Nucleic Acids Res. 2022, 50, 10733–10755. [Google Scholar] [CrossRef]
  197. Cabianca, D.S.; Casa, V.; Bodega, B.; Xynos, A.; Ginelli, E.; Tanaka, Y.; Gabellini, D. A long ncRNA links copy number variation to a polycomb/trithorax epigenetic switch in FSHD muscular dystrophy. Cell 2012, 149, 819–831. [Google Scholar] [CrossRef]
  198. Park, J.; Lee, H.; Han, N.; Kwak, S.; Lee, H.T.; Kim, J.H.; Kang, K.; Youn, B.H.; Yang, J.H.; Jeong, H.J.; et al. Long non-coding RNA ChRO1 facilitates ATRX/DAXX-dependent H3.3 deposition for transcription-associated heterochromatin reorganization. Nucleic Acids Res. 2018, 46, 11759–11775. [Google Scholar] [CrossRef]
  199. Wang, L.; Zhao, Y.; Bao, X.; Zhu, X.; Kwok, Y.K.; Sun, K.; Chen, X.; Huang, Y.; Jauch, R.; Esteban, M.A.; et al. LncRNA Dum interacts with Dnmts to regulate Dppa2 expression during myogenic differentiation and muscle regeneration. Cell Res. 2015, 25, 335–350. [Google Scholar] [CrossRef]
  200. Fan, Y.; Ren, C.; Deng, K.; Zhang, Z.; Li, J.; Deng, M.; Zhang, Y.; Wang, F. The regulation of LncRNA GTL2 expression by DNA methylation during sheep skeletal muscle development. Genomics 2022, 114, 110453. [Google Scholar] [CrossRef]
  201. Cai, B.; Ma, M.; Yuan, R.; Zhou, Z.; Zhang, J.; Kong, S.; Lin, D.; Lian, L.; Li, J.; Zhang, X.; et al. MYH1G-AS is a chromatin-associated lncRNA that regulates skeletal muscle development in chicken. Cell. Mol. Biol. Lett. 2024, 29, 9. [Google Scholar] [CrossRef]
  202. Zheng, Y.; Liu, T.; Li, Q.; Li, J. Integrated analysis of long non-coding RNAs (lncRNAs) and mRNA expression profiles identifies lncRNA PRKG1-AS1 playing important roles in skeletal muscle aging. Aging 2021, 13, 15044–15060. [Google Scholar] [CrossRef]
  203. Ruan, L.; Mendhe, B.; Parker, E.; Kent, A.; Isales, C.M.; Hill, W.D.; McGee-Lawrence, M.; Fulzele, S.; Hamrick, M.W. Long Non-coding RNA MALAT1 Is Depleted with Age in Skeletal Muscle in vivo and MALAT1 Silencing Increases Expression of TGF-beta1 In Vitro. Front. Physiol. 2021, 12, 742004. [Google Scholar] [CrossRef]
  204. Neppl, R.L.; Wu, C.L.; Walsh, K. lncRNA Chronos is an aging-induced inhibitor of muscle hypertrophy. J. Cell Biol. 2017, 216, 3497–3507. [Google Scholar] [CrossRef]
  205. Yu, M.; He, X.; Liu, T.; Li, J. lncRNA GPRC5D-AS1 as a ceRNA inhibits skeletal muscle aging by regulating miR-520d-5p. Aging 2023, 15, 13980–13997. [Google Scholar] [CrossRef]
  206. Li, Y.; Shi, H.; Chen, R.; Zhou, S.; Lei, S.; She, Y. Role of miRNAs and lncRNAs in dexamethasone-induced myotube atrophy In Vitro. Exp. Ther. Med. 2021, 21, 146. [Google Scholar] [CrossRef]
  207. Jin, J.; Du, M.; Wang, J.; Guo, Y.; Zhang, J.; Zuo, H.; Hou, Y.; Wang, S.; Lv, W.; Bai, W.; et al. Conservative analysis of Synaptopodin-2 intron sense-overlapping lncRNA reveals its novel function in promoting muscle atrophy. J. Cachexia Sarcopenia Muscle 2022, 13, 2017–2030. [Google Scholar] [CrossRef]
  208. Li, Z.; Cai, B.; Abdalla, B.A.; Zhu, X.; Zheng, M.; Han, P.; Nie, Q.; Zhang, X. LncIRS1 controls muscle atrophy via sponging miR-15 family to activate IGF1-PI3K/AKT pathway. J. Cachexia Sarcopenia Muscle 2019, 10, 391–410. [Google Scholar] [CrossRef]
  209. Li, J.; Yang, T.; Tang, H.; Sha, Z.; Chen, R.; Chen, L.; Yu, Y.; Rowe, G.C.; Das, S.; Xiao, J. Inhibition of lncRNA MAAT Controls Multiple Types of Muscle Atrophy by cis- and trans-Regulatory Actions. Mol. Ther. 2021, 29, 1102–1119. [Google Scholar] [CrossRef]
  210. Hitachi, K.; Nakatani, M.; Kiyofuji, Y.; Inagaki, H.; Kurahashi, H.; Tsuchida, K. An Analysis of Differentially Expressed Coding and Long Non-Coding RNAs in Multiple Models of Skeletal Muscle Atrophy. Int. J. Mol. Sci. 2021, 22, 2558. [Google Scholar] [CrossRef]
  211. Zhang, Z.K.; Li, J.; Guan, D.; Liang, C.; Zhuo, Z.; Liu, J.; Lu, A.; Zhang, G.; Zhang, B.T. A newly identified lncRNA MAR1 acts as a miR-487b sponge to promote skeletal muscle differentiation and regeneration. J. Cachexia Sarcopenia Muscle 2018, 9, 613–626. [Google Scholar] [CrossRef]
  212. Hitachi, K.; Nakatani, M.; Funasaki, S.; Hijikata, I.; Maekawa, M.; Honda, M.; Tsuchida, K. Expression Levels of Long Non-Coding RNAs Change in Models of Altered Muscle Activity and Muscle Mass. Int. J. Mol. Sci. 2020, 21, 1628. [Google Scholar] [CrossRef]
  213. Zhang, Z.K.; Li, J.; Guan, D.; Liang, C.; Zhuo, Z.; Liu, J.; Lu, A.; Zhang, G.; Zhang, B.T. Long Noncoding RNA lncMUMA Reverses Established Skeletal Muscle Atrophy following Mechanical Unloading. Mol. Ther. 2018, 26, 2669–2680. [Google Scholar] [CrossRef]
  214. Zhang, H.; Wang, F.; Pang, X.; Zhou, Y.; Li, S.; Li, W.; Zhang, P.; Chen, X. Decreased expression of H19/miR-675 ameliorates muscle atrophy by regulating the IGF1R/Akt/FoxO signaling pathway. Mol. Med. 2023, 29, 78. [Google Scholar] [CrossRef]
  215. Lei, S.; She, Y.; Zeng, J.; Chen, R.; Zhou, S.; Shi, H. Expression patterns of regulatory lncRNAs and miRNAs in muscular atrophy models induced by starvation In Vitro and In Vivo. Mol. Med. Rep. 2019, 20, 4175–4185. [Google Scholar] [CrossRef]
  216. Alessio, E.; Buson, L.; Chemello, F.; Peggion, C.; Grespi, F.; Martini, P.; Massimino, M.L.; Pacchioni, B.; Millino, C.; Romualdi, C.; et al. Single cell analysis reveals the involvement of the long non-coding RNA Pvt1 in the modulation of muscle atrophy and mitochondrial network. Nucleic Acids Res. 2019, 47, 1653–1670. [Google Scholar] [CrossRef]
  217. Zhu, M.; Liu, J.; Xiao, J.; Yang, L.; Cai, M.; Shen, H.; Chen, X.; Ma, Y.; Hu, S.; Wang, Z.; et al. Lnc-mg is a long non-coding RNA that promotes myogenesis. Nat. Commun. 2017, 8, 14718. [Google Scholar] [CrossRef]
  218. Hitachi, K.; Nakatani, M.; Takasaki, A.; Ouchi, Y.; Uezumi, A.; Ageta, H.; Inagaki, H.; Kurahashi, H.; Tsuchida, K. Myogenin promoter-associated lncRNA Myoparr is essential for myogenic differentiation. EMBO Rep. 2019, 20, e47468. [Google Scholar] [CrossRef]
  219. Hitachi, K.; Nakatani, M.; Tsuchida, K. Long Non-Coding RNA Myoparr Regulates GDF5 Expression in Denervated Mouse Skeletal Muscle. Noncoding RNA 2019, 5, 33. [Google Scholar] [CrossRef]
  220. Yu, J.; Loh, K.; Yang, H.Q.; Du, M.R.; Wu, Y.X.; Liao, Z.Y.; Guo, A.; Yang, Y.F.; Chen, B.; Zhao, Y.X.; et al. The Whole-transcriptome Landscape of Diabetes-related Sarcopenia Reveals the Specific Function of Novel lncRNA Gm20743. Commun. Biol. 2022, 5, 774. [Google Scholar] [CrossRef]
  221. Chen, R.; Yang, T.; Jin, B.; Xu, W.; Yan, Y.; Wood, N.; Lehmann, H.I.; Wang, S.; Zhu, X.; Yuan, W.; et al. CircTmeff1 Promotes Muscle Atrophy by Interacting with TDP-43 and Encoding A Novel TMEFF1-339aa Protein. Adv. Sci. 2023, 10, e2206732. [Google Scholar] [CrossRef]
  222. Zhao, J.; Zhao, X.; Shen, X.; Zhang, Y.; Zhang, Y.; Ye, L.; Li, D.; Zhu, Q.; Yin, H. CircCCDC91 regulates chicken skeletal muscle development by sponging miR-15 family via activating IGF1-PI3K/AKT signaling pathway. Poult. Sci. 2022, 101, 101803. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The pathogenesis of skeletal muscle atrophy. Skeletal muscle atrophy is often accompanied by activation of the UPS system, impaired autophagic flux, decreased protein synthesis, and elevated levels of oxidative stress, inflammation, caspases and calpain.
Figure 1. The pathogenesis of skeletal muscle atrophy. Skeletal muscle atrophy is often accompanied by activation of the UPS system, impaired autophagic flux, decreased protein synthesis, and elevated levels of oxidative stress, inflammation, caspases and calpain.
Ijms 25 08362 g001
Figure 2. The pathway related to skeletal muscle atrophy. PI3K/AKT, NF-κB, AMPK and myostatin pathways play an important role in muscular atrophy. The PI3K/AKT pathway can promote protein synthesis, inhibit the UPS system and improve autophagy flux. The NF-κB pathway can promote inflammation and the UPS system. AMPK is more responsible, it can inhibit mTOR but promote PGC1α. Myostatin can affect the UPS system through the Smad pathway.
Figure 2. The pathway related to skeletal muscle atrophy. PI3K/AKT, NF-κB, AMPK and myostatin pathways play an important role in muscular atrophy. The PI3K/AKT pathway can promote protein synthesis, inhibit the UPS system and improve autophagy flux. The NF-κB pathway can promote inflammation and the UPS system. AMPK is more responsible, it can inhibit mTOR but promote PGC1α. Myostatin can affect the UPS system through the Smad pathway.
Ijms 25 08362 g002
Table 1. HADCs inhibitors in skeletal muscle atrophy.
Table 1. HADCs inhibitors in skeletal muscle atrophy.
NameMechanismResultsReference
Trichostatin AInhibiting HDAC activityPromoting atrogin-1 mRNA levelAlamdari et al. (2010) [87]
Improving amyotrophic lateral sclerosisYoo et al. (2011) [100]
Improving unloaded-induced skeletal muscle atrophyAucouturier et al. (2015) [101]
Inhibiting HDAC1/2Improving muscle atrophy induced by cigarette smoke exposureDing et al. (2019) [99]
MS-275Inhibiting class I HDACImproving fasting and denervation-induced skeletal muscle atrophyBeharry et al. (2014) [97]
Butyrate Improving sarcopeniaWalsh et al. (2015) [102]
HC toxin Improving starvation-induced muscle atrophyTan et al. (2015) [106]
Valproic acid Attenuating cancer cachexia- induced skeletal muscle atrophySun et al. (2016) [105]
NVS-HD1Inhibiting HDAC4Improving dexamethasone-induced skeletal muscle atrophyLuo et al. (2019) [95]
LMK-235Inhibiting HDAC5Improving unloading-induced muscle atrophyMochalova et al. (2020) [103]
Tubastatin AInhibiting HDAC6Improving Duchenne muscular dystrophyOsseni et al. (2022) [104]
trichostatin A, apicidin, romidepsinInhibiting HDAC1/2attenuating dexamethasone-induced muscle atrophyHabibian et al. (2023) [107]
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

Du, J.; Wu, Q.; Bae, E.J. Epigenetics of Skeletal Muscle Atrophy. Int. J. Mol. Sci. 2024, 25, 8362. https://doi.org/10.3390/ijms25158362

AMA Style

Du J, Wu Q, Bae EJ. Epigenetics of Skeletal Muscle Atrophy. International Journal of Molecular Sciences. 2024; 25(15):8362. https://doi.org/10.3390/ijms25158362

Chicago/Turabian Style

Du, Jiacheng, Qian Wu, and Eun Ju Bae. 2024. "Epigenetics of Skeletal Muscle Atrophy" International Journal of Molecular Sciences 25, no. 15: 8362. https://doi.org/10.3390/ijms25158362

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