Int J Sports Med 2023; 44(11): 763-777
DOI: 10.1055/a-2044-8277
Review

Protein Turnover in Skeletal Muscle: Looking at Molecular Regulation towards an Active Lifestyle

Rita Pinho Ferreira
1   LAQV-REQUIMTE, Department of Chemistry, University of Aveiro, Aveiro, Portugal
,
Jose Alberto Duarte
2   TOXRUN – Toxicology Research Unit, University Institute of Health Sciences, CESPU, CRL, Gandra, Portugal
3   CIAFEL, Faculty of Sports, University of Porto and Laboratory for Integrative and Translational Research in Population Health (ITR), Porto, Portugal
› Author Affiliations
Funding This work was supported by LAQV (UIDB/50006/2020), CIAFEL (UID/DTP/00617/2020) and ITR (LA/P/0064/2020) through national founds by the Portuguese Foundation for Science and Technology (FCT) and co-financed by the European Regional Development Fund (FEDER), within the PT2020 Partnership Agreement.

Abstract

Skeletal muscle is a highly plastic tissue, able to change its mass and functional properties in response to several stimuli. Skeletal muscle mass is influenced by the balance between protein synthesis and breakdown, which is regulated by several signaling pathways. The relative contribution of Akt/mTOR signaling, ubiquitin-proteasome pathway, autophagy among other signaling pathways to protein turnover and, therefore, to skeletal muscle mass, differs depending on the wasting or loading condition and muscle type. By modulating mitochondria biogenesis, PGC-1α has a major role in the cell’s bioenergetic status and, thus, on protein turnover. In fact, rates of protein turnover regulate differently the levels of distinct protein classes in response to atrophic or hypertrophic stimuli. Mitochondrial protein turnover rates may be enhanced in wasting conditions, whereas the increased turnover of myofibrillar proteins triggers muscle mass gain. The present review aims to update the knowledge on the molecular pathways implicated in the regulation of protein turnover in skeletal muscle, focusing on how distinct muscle proteins may be modulated by lifestyle interventions with emphasis on exercise training. The comprehensive analysis of the anabolic effects of exercise programs will pave the way to the tailored management of muscle wasting conditions.



Publication History

Received: 04 November 2022

Accepted: 28 February 2023

Accepted Manuscript online:
28 February 2023

Article published online:
17 July 2023

© 2023. Thieme. All rights reserved.

Georg Thieme Verlag
Rüdigerstraße 14, 70469 Stuttgart, Germany

 
  • References

  • 1 Alldritt I, Greenhaff PL, Wilkinson DJ. Metabolomics as an important tool for determining the mechanisms of human skeletal muscle deconditioning. Int J Mol Sci 2021; 22: 13575
  • 2 Koyama S, Hata S, Witt CC. et al. Muscle RING-Finger protein-1 (MuRF1) as a connector of muscle energy metabolism and protein synthesis. J Mol Biol 2008; 376: 1224-1236
  • 3 Rose AJ, Richter EA. Regulatory mechanisms of skeletal muscle protein turnover during exercise. J Appl Physiol (1985) 2009; 106: 1702-1711
  • 4 Burd NA, De Lisio M. Skeletal muscle remodeling: Interconnections between stem cells and protein turnover. Exerc Sport Sci Rev 2017; 45: 187-191
  • 5 Murach KA, Fry CS, Dupont-Versteegden EE. et al. Fusion and beyond: Satellite cell contributions to loading-induced skeletal muscle adaptation. Faseb J 2021; 35: e21893
  • 6 Pallafacchina G, Blaauw B, Schiaffino S. Role of satellite cells in muscle growth and maintenance of muscle mass. Nutr Metab Cardiovasc Dis 2013; 23: S12-S18
  • 7 Neti G, Novak SM, Thompson VF. et al. Properties of easily releasable myofilaments: are they the first step in myofibrillar protein turnover?. Am J Physiol Cell Physiol 2009; 296: C1383-C1390
  • 8 Phillips SM, McGlory C. CrossTalk proposal: The dominant mechanism causing disuse muscle atrophy is decreased protein synthesis. J Physiol 2014; 592: 5341-5343
  • 9 Attaix D, Mosoni L, Dardevet D. et al. Altered responses in skeletal muscle protein turnover during aging in anabolic and catabolic periods. Int J Biochem Cell Biol 2005; 37: 1962-1973
  • 10 van der Westhuyzen DR, Matsumoto K, Etlinger JD. Easily releasable myofilaments from skeletal and cardiac muscles maintained in vitro. Role in myofibrillar assembly and turnover. J Biol Chem 1981; 256: 11791-11797
  • 11 Papageorgopoulos C, Caldwell K, Schweingrubber H. et al. Measuring synthesis rates of muscle creatine kinase and myosin with stable isotopes and mass spectrometry. Anal Biochem 2002; 309: 1-10
  • 12 Terjung RL. Cytochrome c turnover in skeletal muscle. Biochem Biophys Res Commun 1975; 66: 173-178
  • 13 Rucklidge GJ, Milne G, McGaw BA. et al. Turnover rates of different collagen types measured by isotope ratio mass spectrometry. Biochim Biophys Acta 1992; 1156: 57-61
  • 14 Short KR, Nair KS. Muscle protein metabolism and the sarcopenia of aging. Int J Sport Nutr Exerc Metab 2001; 11: S119-S127
  • 15 Kimball SR, O’Malley JP, Anthony JC. et al. Assessment of biomarkers of protein anabolism in skeletal muscle during the life span of the rat: sarcopenia despite elevated protein synthesis. Am J Physiol Endocrinol Metab 2004; 287: E772-E780
  • 16 Miller BF, Baehr LM, Musci RV. et al. Muscle-specific changes in protein synthesis with aging and reloading after disuse atrophy. J Cachexia Sarcopenia Muscle 2019; 10: 1195-1209
  • 17 Mobley CB, Holland AM, Kephart WC. et al. Progressive resistance-loaded voluntary wheel running increases hypertrophy and differentially affects muscle protein synthesis, ribosome biogenesis, and proteolytic markers in rat muscle. J Anim Physiol Anim Nutr 2018; 102: 317-329
  • 18 van Wessel T, de Haan A, van der Laarse WJ. et al. The muscle fiber type-fiber size paradox: hypertrophy or oxidative metabolism?. Eur J Appl Physiol 2010; 110: 665-694
  • 19 Rasmussen BB, Richter EA. The balancing act between the cellular processes of protein synthesis and breakdown: exercise as a model to understand the molecular mechanisms regulating muscle mass. J Appl Physiol (1985) 2009; 106: 1365-1366
  • 20 von Walden F, Liu C, Aurigemma N. et al. mTOR signaling regulates myotube hypertrophy by modulating protein synthesis, rDNA transcription, and chromatin remodeling. Am J Physiol Cell Physiol 2016; 311: C663-C672
  • 21 Armstrong DD, Esser KA. Wnt/beta-catenin signaling activates growth-control genes during overload-induced skeletal muscle hypertrophy. Am J Physiol Cell Physiol 2005; 289: C853-C859
  • 22 Fujimaki S, Matsumoto T, Muramatsu M. et al. The endothelial Dll4-muscular Notch2 axis regulates skeletal muscle mass. Nat Metab 2022; 4: 180-189
  • 23 Mavropalias G, Sim M, Taaffe DR. et al. Exercise medicine for cancer cachexia: targeted exercise to counteract mechanisms and treatment side effects. J Cancer Res Clin Oncol 2022; 148: 1389-1406
  • 24 Blaauw B, Schiaffino S, Reggiani C. Mechanisms modulating skeletal muscle phenotype. Compr Physiol 2013; 3: 1645-1687
  • 25 Boppart MD, Mahmassani ZS. Integrin signaling: linking mechanical stimulation to skeletal muscle hypertrophy. Am J Physiol Cell Physiol 2019; 317: C629-C641
  • 26 Koskinen SO, Kjaer M, Mohr T. et al. Type IV collagen and its degradation in paralyzed human muscle: effect of functional electrical stimulation. Muscle Nerve 2000; 23: 580-589
  • 27 Gharahdaghi N, Phillips BE, Szewczyk NJ. et al. Links between testosterone, oestrogen, and the growth hormone/insulin-like growth factor axis and resistance exercise muscle adaptations. Front Physiol 2021; 11: 621226
  • 28 Schakman O, Kalista S, Barbé C. et al. Glucocorticoid-induced skeletal muscle atrophy. Int J Biochem Cell Biol 2013; 45: 2163-2172
  • 29 Pedersen BK, Steensberg A, Schjerling P. Muscle-derived interleukin-6: possible biological effects. J Physiol 2001; 536: 329-337
  • 30 McGlory C, Devries MC, Phillips SM. Skeletal muscle and resistance exercise training; the role of protein synthesis in recovery and remodeling. J Appl Physiol (1985) 2017; 122: 541-548
  • 31 Merle A, Jollet M, Britto FA. et al. Endurance exercise decreases protein synthesis and ER-mitochondria contacts in mouse skeletal muscle. J Appl Physiol (1985) 2019; 127: 1297-1306
  • 32 Chaillou T. Ribosome specialization and its potential role in the control of protein translation and skeletal muscle size. J Appl Physiol (1985) 2019; 127: 599-607
  • 33 Nakada S, Ogasawara R, Kawada S. et al. Correlation between ribosome biogenesis and the magnitude of hypertrophy in overloaded skeletal muscle. PLoS One 2016; 11: e0147284
  • 34 Stec MJ, Kelly NA, Many GM. et al. Ribosome biogenesis may augment resistance training-induced myofiber hypertrophy and is required for myotube growth in vitro. Am J Physiol Endocrinol Metab 2016; 310: E652-E661
  • 35 Figueiredo VC, Caldow MK, Massie V. et al. Ribosome biogenesis adaptation in resistance training-induced human skeletal muscle hypertrophy. Am J Physiol Endocrinol Metab 2015; 309: E72-E83
  • 36 Perks KL, Rossetti G, Kuznetsova I. et al. PTCD1 is required for 16S rRNA maturation complex stability and mitochondrial ribosome assembly. Cell Rep 2018; 23: 127-142
  • 37 von Walden F. Ribosome biogenesis in skeletal muscle: coordination of transcription and translation. J Appl Physiol (1985) 2019; 127: 591-598
  • 38 Smiles WJ, Hawley JA, Camera DM. Effects of skeletal muscle energy availability on protein turnover responses to exercise. J Exp Biol 2016; 219: 214-225
  • 39 Reynolds TH, Bodine SC, Lawrence JCL. Control of Ser2448 phosphorylation in the mammalian target of rapamycin by insulin and skeletal muscle load. J Biol Chem 2002; 277: 17657-17662
  • 40 Rommel C, Bodine SC, Clarke BA. et al. Mediation of IGF-1-induced skeletal myotube hypertrophy by PI(3)K/Akt/mTOR and PI(3)K/Akt/GSK3 pathways. Nat Cell Biol 2001; 3: 1009-1013
  • 41 Guridi M, Kupr B, Romanino K. et al. Alterations to mTORC1 signaling in the skeletal muscle differentially affect whole-body metabolism. Skelet Muscle 2016; 6: 13
  • 42 Belova SP, Vilchinskaya NA, Mochalova EP. et al. Elevated p70S6K phosphorylation in rat soleus muscle during the early stage of unloading: Causes and consequences. Arch Biochem Biophys 2019; 674: 108105
  • 43 Wakatsuki T, Ohira Y, Yasui W. et al. Responses of contractile properties in rat soleus to high-energy phosphates and/or unloading. Jpn J Physiol 1994; 44: 193-204
  • 44 Ross FA, MacKintosh C, Hardie DG. AMP-activated protein kinase: a cellular energy sensor that comes in 12 flavours. FEBS J 2016; 283: 2987-3001
  • 45 Drake JC, Wilson RJ, Laker RC. et al. Mitochondria-localized AMPK responds to local energetics and contributes to exercise and energetic stress-induced mitophagy. Proc Natl Acad Sci USA 2021; 118: e2025932118
  • 46 Murakami T, Hasegawa K, Yoshinaga M. Rapid induction of REDD1 expression by endurance exercise in rat skeletal muscle. Biochem Biophys Res Commun 2011; 405: 615-619
  • 47 Bolster DR, Crozier SJ, Kimball SR. et al. AMP-activated protein kinase suppresses protein synthesis in rat skeletal muscle through down-regulated mammalian target of rapamycin (mTOR) signaling. J Biol Chem 2002; 277: 23977-23980
  • 48 Mirzoev TM. Skeletal muscle recovery from disuse atrophy: protein turnover signaling and strategies for accelerating muscle regrowth. Int J Mol Sci 2020; 21: 7940
  • 49 Britto FA, Dumas K, Giorgetti-Peraldi S. et al. Is REDD1 a metabolic double agent? Lessons from physiology and pathology. Am J Physiol Cell Physiol 2020; 319: C807-C824
  • 50 White JP, Puppa MJ, Gao S. et al. Muscle mTORC1 suppression by IL-6 during cancer cachexia: a role for AMPK. Am J Physiol Endocrinol Metab 2013; 304: E1042-E1052
  • 51 Hain BA, Xu H, VanCleave AM. et al. REDD1 deletion attenuates cancer cachexia in mice. J Appl Physiol (1985) 2021; 131: 1718-1730
  • 52 Jee H, Chang J-E, Yang EJ. Positive prehabilitative effect of intense treadmill exercise for ameliorating cancer cachexia symptoms in a mouse model. J Cancer 2016; 7: 2378-2387
  • 53 Padrão AI, Figueira ACC, Faustino-Rocha AI. et al. Long-term exercise training prevents mammary tumorigenesis-induced muscle wasting in rats through the regulation of TWEAK signalling. Acta Physiol (Oxf) 2017; 219: 803-813
  • 54 Smiles WJ, Hawley JA, Camera DM. Effects of skeletal muscle energy availability on protein turnover responses to exercise. J Exp Biol 2016; 219: 214-225
  • 55 Hulmi JJ, Oliveira BM, Silvennoinen M. et al. Muscle protein synthesis, mTORC1/MAPK/Hippo signaling, and capillary density are altered by blocking of myostatin and activins. Am J Physiol Endocrinol Metab 2013; 304: E41-E50
  • 56 Rodriguez J, Vernus B, Chelh I. et al. Myostatin and the skeletal muscle atrophy and hypertrophy signaling pathways. Cell Mol Life Sci 2014; 71: 4361-4371
  • 57 Trendelenburg AU, Meyer A, Rohner D. et al. Myostatin reduces Akt/TORC1/p70S6K signaling, inhibiting myoblast differentiation and myotube size. Am J Physiol Cell Physiol 2009; 296: C1258-C1270
  • 58 Sharma M, McFarlane C, Kambadur R. et al. Myostatin: expanding horizons. IUBMB Life 2015; 67: 589-600
  • 59 Lokireddy S, Mouly V, Butler-Browne G. et al. Myostatin promotes the wasting of human myoblast cultures through promoting ubiquitin-proteasome pathway-mediated loss of sarcomeric proteins. Am J Physiol Cell Physiol 2011; 301: C1316-C1324
  • 60 Durieux AC, Amirouche A, Banzet S. et al. Ectopic expression of myostatin induces atrophy of adult skeletal muscle by decreasing muscle gene expression. Endocrinology 2007; 148: 3140-3147
  • 61 Tang L, Cao W, Zhao T. et al. Weight-bearing exercise prevents skeletal muscle atrophy in ovariectomized rat. J Physiol Biochem 2021; 77: 273-281
  • 62 Wu Y, Collier L, Qin W. et al. Electrical stimulation modulates Wnt signaling and regulates genes for the motor endplate and calcium binding in muscle of rats with spinal cord transection. BMC Neurosci 2013; 14: 81
  • 63 Gnimassou O, Francaux M, Deldicque L. Hippo pathway and skeletal muscle mass regulation in mammals: a controversial relationship. Front Physiol 2017; 8: 190
  • 64 Watt KI, Goodman CA, Hornberger TA. et al. The Hippo signaling pathway in the regulation of skeletal muscle mass and function. Exerc Sport Sci Rev 2018; 46: 92-96
  • 65 Watt KI, Turner BJ, Hagg A. et al. The Hippo pathway effector YAP is a critical regulator of skeletal muscle fibre size. Nat Commun 2015; 6: 6048
  • 66 Goodman CA, Dietz JM, Jacobs BL. et al. Yes-Associated Protein is up-regulated by mechanical overload and is sufficient to induce skeletal muscle hypertrophy. FEBS Lett 2015; 589: 1491-1497
  • 67 Judson RN, Gray SR, Walker C. et al. Constitutive expression of Yes-associated protein (Yap) in adult skeletal muscle fibres induces muscle atrophy and myopathy. PLoS One 2013; 8: e59622
  • 68 Liu D, Sartor MA, Nader GA. et al. Skeletal muscle gene expression in response to resistance exercise: sex specific regulation. BMC Genomics 2010; 11: 659
  • 69 Kumar V, Atherton P, Smith K. et al. Human muscle protein synthesis and breakdown during and after exercise. J Appl Physiol (1985) 2009; 106: 2026-2039
  • 70 Costelli P, Reffo P, Penna F. et al. Ca(2+)-dependent proteolysis in muscle wasting. Int J Biochem Cell Biol 2005; 37: 2134-2146
  • 71 Bechet D, Tassa A, Taillandier D. et al. Lysosomal proteolysis in skeletal muscle. Int J Biochem Cell Biol 2005; 37: 2098-2114
  • 72 Pasiakos SM, Carbone JW. Assessment of skeletal muscle proteolysis and the regulatory response to nutrition and exercise. IUBMB Life 2014; 66: 478-484
  • 73 Christian BE, Spremulli LL. Mechanism of protein biosynthesis in mammalian mitochondria. Biochim Biophys Acta 2012; 1819: 1035-1054
  • 74 Cunha TF, Moreira JBN, Paixão NA. et al. Aerobic exercise training upregulates skeletal muscle calpain and ubiquitin-proteasome systems in healthy mice. J Appl Physiol (1985) 2012; 112: 1839-1846
  • 75 Nixon RA. The calpains in aging and aging-related diseases. Ageing Res Rev 2003; 2: 407-418
  • 76 Dayton WR, Reville WJ, Goll DE. et al. A Ca2+-activated protease possibly involved in myofibrillar protein turnover. Partial characterization of the purified enzyme. Biochemistry 1976; 15: 2159-2167
  • 77 Kumamoto T, Kleese WC, Cong JY. et al. Localization of the Ca(2+)-dependent proteinases and their inhibitor in normal, fasted, and denervated rat skeletal muscle. Anat Rec 1992; 232: 60-77
  • 78 Otani K, Han D-H, Ford EL. et al. Calpain system regulates muscle mass and glucose transporter GLUT4 turnover. J Biol Chem 2004; 279: 20915-20920
  • 79 Beckmann JS, Spencer M. Calpain 3, the “gatekeeper” of proper sarcomere assembly, turnover and maintenance. Neuromuscul Disord 2008; 18: 913-921
  • 80 Ojima K, Ono Y, Doi N. et al. Myogenic stage, sarcomere length, and protease activity modulate localization of muscle-specific calpain. J Biol Chem 2007; 282: 14493-14504
  • 81 Salazar JJ, Michele DE, Brooks SV. Inhibition of calpain prevents muscle weakness and disruption of sarcomere structure during hindlimb suspension. J Appl Physiol (1985) 2010; 108: 120-127
  • 82 Enns DL, Belcastro AN. Early activation and redistribution of calpain activity in skeletal muscle during hindlimb unweighting and reweighting. Can J Physiol Pharmacol 2006; 84: 601-609
  • 83 Murphy RM, Goodman CA, McKenna MJ. et al. Calpain-3 is autolyzed and hence activated in human skeletal muscle 24 h following a single bout of eccentric exercise. J Appl Physiol (1985) 2007; 103: 926-931
  • 84 Murphy RM, Lamb GD. Endogenous calpain-3 activation is primarily governed by small increases in resting cytoplasmic [Ca2+] and is not dependent on stretch. J Biol Chem 2009; 284: 7811-7819
  • 85 Zhang B-T, Whitehead NP, Gervasio OL. et al. Pathways of Ca2+entry and cytoskeletal damage following eccentric contractions in mouse skeletal muscle. J Appl Physiol (1985) 2012; 112: 2077-2086
  • 86 Raastad T, Owe SG, Paulsen G. et al. Changes in calpain activity, muscle structure, and function after eccentric exercise. Med Sci Sports Exerc 2010; 42: 86-95
  • 87 Sacheck JM, Hyatt J-PK, Raffaello A. et al. Rapid disuse and denervation atrophy involve transcriptional changes similar to those of muscle wasting during systemic diseases. FASEB J 2007; 21: 140-155
  • 88 Okamoto T, Torii S, Machida S. Differential gene expression of muscle-specific ubiquitin ligase MAFbx/Atrogin-1 and MuRF1 in response to immobilization-induced atrophy of slow-twitch and fast-twitch muscles. J Physiol Sci 2011; 61: 537-546
  • 89 Bodine SC, Edward F. Adolph Distinguished Lecture. Skeletal muscle atrophy: Multiple pathways leading to a common outcome. J Appl Physiol (1985) 2020; 129: 272-282
  • 90 Choi RH, McConahay A, Jeong H-W. et al. Tribbles 3 regulates protein turnover in mouse skeletal muscle. Biochem Biophys Res Commun 2017; 493: 1236-1242
  • 91 An D, Lessard SJ, Toyoda T. et al. Overexpression of TRB3 in muscle alters muscle fiber type and improves exercise capacity in mice. Am J Physiol Regul Integr Comp Physiol 2014; 306: R925-R933
  • 92 Sato AY, Richardson D, Cregor M. et al. Glucocorticoids induce bone and muscle atrophy by tissue-specific mechanisms upstream of E3 ubiquitin ligases. Endocrinology 2017; 158: 664-677
  • 93 Milan G, Romanello V, Pescatore F. et al. Regulation of autophagy and the ubiquitin-proteasome system by the FoxO transcriptional network during muscle atrophy. Nat Commun 2015; 6: 6670
  • 94 Witt SH, Granzier H, Witt CC. et al. MURF-1 and MURF-2 target a specific subset of myofibrillar proteins redundantly: towards understanding MURF-dependent muscle ubiquitination. J Mol Biol 2005; 350: 713-722
  • 95 Mei Z, Zhang D, Hu B. et al. FBXO32 targets c-myc for proteasomal degradation and inhibits c-myc activity. J Biol Chem 2015; 290: 16202-16214
  • 96 Lagirand-Cantaloube J, Offner N, Csibi A. et al. The initiation factor eIF3-f is a major target for atrogin1/MAFbx function in skeletal muscle atrophy. EMBO J 2008; 27: 1266-1276
  • 97 McElhinny AS, Kakinuma K, Sorimachi H. et al. Muscle-specific RING finger-1 interacts with titin to regulate sarcomeric M-line and thick filament structure and may have nuclear functions via its interaction with glucocorticoid modulatory element binding protein-1. J Cell Biol 2002; 157: 125-136
  • 98 Romanello V, Sandri M. The connection between the dynamic remodeling of the mitochondrial network and the regulation of muscle mass. Cell Mol Life Sci 2021; 78: 1305-1328
  • 99 Romanello V, Sandri M. Mitochondrial quality control and muscle mass maintenance. Front Physiol 2016; 6: 422
  • 100 Bentzinger CF, Romanino K, Cloëtta D. et al. Skeletal muscle-specific ablation of raptor, but not of rictor, causes metabolic changes and results in muscle dystrophy. Cell Metab 2008; 8: 411-424
  • 101 Morita M, Gravel S-P, Chénard V. et al. mTORC1 controls mitochondrial activity and biogenesis through 4E-BP-dependent translational regulation. Cell Metab 2013; 18: 698-711
  • 102 Andreux PA, Houtkooper RH, Auwerx J. Pharmacological approaches to restore mitochondrial function. Nat Rev Drug Discov 2013; 12: 465-483
  • 103 Cunningham JT, Rodgers JT, Arlow DH. et al. mTOR controls mitochondrial oxidative function through a YY1-PGC-1alpha transcriptional complex. Nature 2007; 450: 736-740
  • 104 Mammucari C, Gherardi G, Zamparo I. et al. The mitochondrial calcium uniporter controls skeletal muscle trophism in vivo. Cell Rep 2015; 10: 1269-1279
  • 105 Stefani DD, Raffaello A, Teardo E. et al. A forty-kilodalton protein of the inner membrane is the mitochondrial calcium uniporter. Nature 2011; 476: 336-340
  • 106 Ruas JL, White JP, Rao RR. et al. A PGC-1α isoform induced by resistance training regulates skeletal muscle hypertrophy. Cell 2012; 151: 1319-1331
  • 107 Arany Z. PGC-1 coactivators and skeletal muscle adaptations in health and disease. Curr Opin Genet Dev 2008; 18: 426-434
  • 108 Jäger S, Handschin C, St-Pierre J. et al. AMP-activated protein kinase (AMPK) action in skeletal muscle via direct phosphorylation of PGC-1alpha. Proc Natl Acad Sci USA 2007; 104: 12017-12022
  • 109 Yokokawa T, Kido K, Suga T. et al. Exercise-induced mitochondrial biogenesis coincides with the expression of mitochondrial translation factors in murine skeletal muscle. Physiol Rep 2018; 6: e13893
  • 110 Yokokawa T, Mori R, Suga T. et al. Muscle denervation reduces mitochondrial biogenesis and mitochondrial translation factor expression in mice. Biochem Biophys Res Commun 2020; 527: 146-152
  • 111 Egan B, Zierath JR. Exercise metabolism and the molecular regulation of skeletal muscle adaptation. Cell Metab 2013; 17: 162-184
  • 112 Anand R, Langer T, Baker MJ. Proteolytic control of mitochondrial function and morphogenesis. Biochim Biophys Acta 2013; 1833: 195-204
  • 113 Padrão AI, Carvalho T, Vitorino R. et al. Impaired protein quality control system underlies mitochondrial dysfunction in skeletal muscle of streptozotocin-induced diabetic rats. Biochim Biophys Acta 2012; 1822: 1189-1197
  • 114 Padrão AI, Oliveira P, Vitorino R. et al. Bladder cancer-induced skeletal muscle wasting: disclosing the role of mitochondria plasticity. Int J Biochem Cell Biol 2013; 45: 1399-1409
  • 115 Drake JC, Yan Z. Mitophagy in maintaining skeletal muscle mitochondrial proteostasis and metabolic health with ageing. J Physiol 2017; 20: 6391-6399
  • 116 Laker RC, Drake JC, Wilson RJ. et al. AMPK phosphorylation of Ulk1 is required for targeting of mitochondria to lysosomes in exercise-induced mitophagy. Nat Commun 2017; 8: 548
  • 117 Vainshtein A, Tryon LD, Pauly M. et al. Role of PGC-1α during acute exercise-induced autophagy and mitophagy in skeletal muscle. Am J Physiol Cell Physiol 2015; 308: C710-C719
  • 118 Guan Y, Drake JC, Yan Z. Exercise-induced mitophagy in skeletal muscle and heart. Exerc Sport Sci Rev 2019; 47: 151-156
  • 119 Yang X, Xue P, Chen H. et al. Denervation drives skeletal muscle atrophy and induces mitochondrial dysfunction, mitophagy and apoptosis via miR-142a-5p/MFN1 axis. Theranostics 2020; 10: 1415-1432
  • 120 Lira VA, Okutsu M, Zhang M. et al. Autophagy is required for exercise training-induced skeletal muscle adaptation and improvement of physical performance. FASEB J 2013; 27: 4184-4193
  • 121 Gonzalez-Freire M, Adelnia F, Moaddel R. et al. Searching for a mitochondrial root to the decline in muscle function with ageing. J Cachexia Sarcopenia Muscle 2018; 9: 435-440
  • 122 Csapo R, Gumpenberger M, Wessner B. Skeletal muscle extracellular matrix - what do we know about its composition, regulation, and physiological roles? A narrative review. Front Physiol 2020; 11: 253
  • 123 Kjaer M, Magnusson P, Krogsgaard M. et al. Extracellular matrix adaptation of tendon and skeletal muscle to exercise. J Anat 2006; 208: 445-450
  • 124 Kritikaki E, Asterling R, Ward L. et al. Exercise training-induced extracellular matrix protein adaptation in locomotor muscles: a systematic review. Cells 2021; 10: 1022
  • 125 Takala TE, Virtanen P. Biochemical composition of muscle extracellular matrix: the effect of loading. Scand J Med Sci Sports 2000; 10: 321-325
  • 126 Lo Presti R, Hopps E, Caimi G. Gelatinases and physical exercise: A systematic review of evidence from human studies. Medicine (Baltimore) 2017; 96: e8072
  • 127 Booth FW, Ruegsegger GN, Toedebusch RG. et al. Endurance exercise activates matrix metalloproteinases in human skeletal muscle. Prog Mol Biol Transl Sci 2015; 135: 129-151
  • 128 Heinemeier KM, Olesen JL, Haddad F. et al. Expression of collagen and related growth factors in rat tendon and skeletal muscle in response to specific contraction types. J Physiol 2007; 582: 1303-1316
  • 129 Moore DR, Phillips SM, Babraj JA. et al. Myofibrillar and collagen protein synthesis in human skeletal muscle in young men after maximal shortening and lengthening contractions. Am J Physiol Endocrinol Metab 2005; 288: E1153-E1159
  • 130 Koskinen SOA, Ahtikoski AM, Komulainen J. et al. Short-term effects of forced eccentric contractions on collagen synthesis and degradation in rat skeletal muscle. Pflugers Arch 2002; 444: 59-72
  • 131 Haus JM, Carrithers JA, Trappe SW. et al. Collagen, cross-linking, and advanced glycation end products in aging human skeletal muscle. J Appl Physiol (1985) 2007; 103: 2068-2076
  • 132 Järvinen TAH, Józsa L, Kannus P. et al. Organization and distribution of intramuscular connective tissue in normal and immobilized skeletal muscles. An immunohistochemical, polarization and scanning electron microscopic study. J Muscle Res Cell Motil 2002; 23: 245-254
  • 133 Mahdy MAA. Skeletal muscle fibrosis: an overview. Cell Tissue Res 2019; 375: 575-588
  • 134 Holeček M, Mičuda S. Amino acid concentrations and protein metabolism of two types of rat skeletal muscle in postprandial state and after brief starvation. Physiol Res 2017; 66: 959-967
  • 135 Muthny T, Kovarik M, Sispera L. et al. The effect of new proteasome inhibitors, belactosin A and C, on protein metabolism in isolated rat skeletal muscle. J Physiol Biochem 2009; 65: 137-146
  • 136 Muthny T, Kovarik M, Sispera L. et al. Protein metabolism in slow- and fast-twitch skeletal muscle during turpentine-induced inflammation. Int J Exp Pathol 2008; 89: 64-71
  • 137 Holeček M. The role of skeletal muscle in the pathogenesis of altered concentrations of branched-chain amino acids (valine, leucine, and isoleucine) in liver cirrhosis, diabetes, and other diseases. Physiol Res 2021; 70: 293-305
  • 138 Sancak Y, Bar-Peled L, Zoncu R. et al. Ragulator-Rag complex targets mTORC1 to the lysosomal surface and is necessary for its activation by amino acids. Cell 2010; 141: 290-303
  • 139 Han JM, Jeong SJ, Park MC. et al. Leucyl-tRNA synthetase is an intracellular leucine sensor for the mTORC1-Signaling Pathway. Cell 2012; 149: 410-424
  • 140 Peng M, Yin N, Li MO. Sestrins function as guanine nucleotide dissociation inhibitors for Rag GTPases to control mTORC1 signaling. Cell 2014; 159: 122-133
  • 141 Zhao Y, Cholewa J, Shang H. et al. Exercise may promote skeletal muscle hypertrophy via enhancing leucine-sensing: preliminary evidence. Front Physiol 2021; 12: 741038
  • 142 Xu D, Shimkus KL, Lacko HA. et al. Evidence for a role for Sestrin1 in mediating leucine-induced activation of mTORC1 in skeletal muscle. Am J Physiol Endocrinol Metab 2019; 316: E817-E828
  • 143 Shimkus KL, Jefferson LS, Gordon BS. et al. Repressors of mTORC1 act to blunt the anabolic response to feeding in the soleus muscle of a cast-immobilized mouse hindlimb. Physiol Rep 2018; 6: e13891
  • 144 Zeng N, D’Souza RF, Figueiredo VC. et al. Acute resistance exercise induces Sestrin2 phosphorylation and p62 dephosphorylation in human skeletal muscle. Physiol Rep 2017; 5: e13526
  • 145 Crisol BM, Lenhare L, Gaspar RS. et al. The role of physical exercise on Sestrin1 and 2 accumulations in the skeletal muscle of mice. Life Sci 2018; 194: 98-103
  • 146 Kamei Y, Hatazawa Y, Uchitomi R. et al. Regulation of skeletal muscle function by amino acids. Nutrients 2020; 12: 261
  • 147 Hatazawa Y, Tadaishi M, Nagaike Y. et al. PGC-1α-mediated branched-chain amino acid metabolism in the skeletal muscle. PLoS One 2014; 9: e91006
  • 148 Hatazawa Y, Qian K, Gong D-W. et al. PGC-1α regulates alanine metabolism in muscle cells. PLoS One 2018; 13: e0190904
  • 149 Hatazawa Y, Senoo N, Tadaishi M. et al. Metabolomic analysis of the skeletal muscle of mice overexpressing PGC-1α. PLoS One 2015; 10: e0129084
  • 150 Macnaughton LS, Wardle SL, Witard OC. et al. The response of muscle protein synthesis following whole-body resistance exercise is greater following 40 g than 20 g of ingested whey protein. Physiol Rep 2016; 4: e12893
  • 151 Estoche JM, Jacinto JL, Roveratti MC. et al. Branched-chain amino acids do not improve muscle recovery from resistance exercise in untrained young adults. Amino Acids 2019; 51: 1387-1395
  • 152 Phillips SM, Glover EI, Rennie MJ. Alterations of protein turnover underlying disuse atrophy in human skeletal muscle. J Appl Physiol (1985) 2009; 107: 645-654
  • 153 Bland KA, Kouw IWK, van Loon LJC. et al. Exercise-based interventions to counteract skeletal muscle mass loss in people with cancer: can we overcome the odds?. Sports Med 2022; 52: 1009-1027