An Evidence-Based Narrative Review of Mechanisms of Resistance Exercise-Induced Human Skeletal Muscle Hypertrophy

Changhyun Lim, Everson A Nunes, Brad S Currier, Jonathan C McLeod, Aaron C Q Thomas, Stuart M Phillips, Changhyun Lim, Everson A Nunes, Brad S Currier, Jonathan C McLeod, Aaron C Q Thomas, Stuart M Phillips

Abstract

Skeletal muscle plays a critical role in physical function and metabolic health. Muscle is a highly adaptable tissue that responds to resistance exercise (RE; loading) by hypertrophying, or during muscle disuse, RE mitigates muscle loss. Resistance exercise training (RET)-induced skeletal muscle hypertrophy is a product of external (e.g., RE programming, diet, some supplements) and internal variables (e.g., mechanotransduction, ribosomes, gene expression, satellite cells activity). RE is undeniably the most potent nonpharmacological external variable to stimulate the activation/suppression of internal variables linked to muscular hypertrophy or countering disuse-induced muscle loss. Here, we posit that despite considerable research on the impact of external variables on RET and hypertrophy, internal variables (i.e., inherent skeletal muscle biology) are dominant in regulating the extent of hypertrophy in response to external stimuli. Thus, identifying the key internal skeletal muscle-derived variables that mediate the translation of external RE variables will be pivotal to determining the most effective strategies for skeletal muscle hypertrophy in healthy persons. Such work will aid in enhancing function in clinical populations, slowing functional decline, and promoting physical mobility. We provide up-to-date, evidence-based perspectives of the mechanisms regulating RET-induced skeletal muscle hypertrophy.

Copyright © 2022 The Author(s). Published by Wolters Kluwer Health, Inc. on behalf of the American College of Sports Medicine.

Figures

FIGURE 1
FIGURE 1
The formulation of external vs internal variables for skeletal muscle hypertrophy. Although external variables (input) are indispensable to activate internal variables for skeletal muscle hypertrophy (output), the response of internal variables stimulated by external variables is the key determinant of skeletal muscle hypertrophy. The size of the color squares reflects the extent of the variables’ contribution to muscle hypertrophy. We are uncertain of the nature of the interaction between the variables and have shown them as multiplicative (•); however, we acknowledge that the function could be additive or more complex (i.e., an unknown function) relationship.
FIGURE 2
FIGURE 2
Overview of mTORC1-related protein synthesis pathways. Multiple signaling cascades converge on mTORC1 and contribute to protein translation. Cytosolic leucine and arginine relieve inhibition on GATOR2 by Sestrin2 and CASTOR1, respectively, and subsequently promote mTORC1-induced protein synthesis via GATOR2–GATOR1–Rag signaling. Mechanical stress causes Cγ1 to colocalize around FAK, catalyzing Ptdlns(4,5)P2 conversion to PA, and subsequently activates the HIPPO pathway (YAP and TAZ). Filamin-C and Bag3 interaction also activates the HIPPO pathway. The HIPPO pathway promotes mTORC1 signaling via the suppression of PTEN. The PI3K–Akt–TSC signaling cascade regulates mTORC1 signaling via Rheb, although Akt can directly promote mTORC1 signaling via PRAS40. AMPK and REDD1 inhibit mTORC1 via TSC; AMPK can also inhibit mTORC1 directly via Raptor. mTORC1 promotes protein synthesis by activating several translational and ribosomal components, and Ras–MEK–ERK 1/2 signaling converges on common factors. 4EBP-1, eukaryotic translation initiation factor 4E-binding protein 1; Akt, protein kinase B; AMPK, AMP-activated protein kinase; Arg, arginine; eEF2, eukaryotic elongation factor 2; eIF4B, eukaryotic translation initiation factor 4B; eIF4E, eukaryotic translation initiation factor 4E; eIF4G, eukaryotic translation initiation factor 4G; ERK 1/2, extracellular signal-related kinase 1/2; FAK, focal adhesion kinase; IGF-1, insulin-like growth factor 1; Leu, leucine; MEK, mitogen-activated protein kinase; MNK1, MAP kinase-interacting kinase 1; mTORC1, mechanistic target of rapamycin complex 1; p70S6K, p70 S6 kinase 1; p90RSK, p90 ribosomal protein S6 kinase; PI3K, phosphoinositide 3-kinase; PRAS40, proline-rich AKT substrate 40 kDa; PtdIns(4,5)P2, phosphatidylinositol 4,5-biphosphate; PTEN, phosphatase and tensin homolog; Raptor, regulatory-associated protein of mTOR; REDD1, regulated in development and DNA damage responses 1; Rheb, Ras homolog enriched in brain; rpS6, ribosomal Protein S6; TAZ, transcriptional coactivator with PDZ-binding motif; TSC, tuberous sclerosis complex; YAP, Yes-associated protein 1.
FIGURE 3
FIGURE 3
A proposed framework of changes in MPS, translational capacity, and muscle fiber CSA in response to RET. The overarching concept is that initial increases in MPS are a biological response to support remodeling of damaged muscle protein and eventually muscle hypertrophy. A, The early-stage increases in MPS are sustained partly by a concomitant elevated translational capacity to support the remodeling of damaged structural and contractile elements of the muscle proteome. B, After the attenuation of exercise-induced muscle damage, there is a reduction in the contribution of MPS to the remodeling of proteins related to the structural and architectural apparatus toward contractile muscle proteins. C, After a relatively short period of time, the rates of MPS are subsequently regulated by the adaptive increase in translational efficiency. D, The result is a detectable increase in skeletal muscle size and mass. All these responses are designed to support an expansion of the muscle protein pool, that is, fiber CSA. The schematic and legend are adapted (with permission) from McGlory et al. (53). A.U., arbitrary units.
FIGURE 4
FIGURE 4
A comparison of the changes in exogenous vs endogenous testosterone in serum. A, Schematic of the circadian changes of serum testosterone throughout 11 d showing the effect of one bout of RE in healthy young men (RE-induced ~30 min testosterone peak; based on data from Willoughby and Taylor [87]). In addition, a representation of the effect of an intramuscular testosterone injection (200 mg testosterone enanthate) vs a schematic of the normal diurnal variation in testosterone concentrations throughout a given week (based on data from Dobs et al. (88)). B, Cumulative AUC of serum testosterone over the first 7 d comparing 1) DV changes, 2) DV + 1 bouts of RE, and 3) 200 mg T enanthate. The schematic is adapted (with permission) from Schroeder et al. (86). AUC, area under the curve; DV, diurnal variation; T, testosterone.
FIGURE 5
FIGURE 5
Representative factors that contribute to the anabolic resistance of skeletal muscle with aging and disease. This schematic is adapted (with permission) from McKendry et al. (2). E, estrogen; T, testosterone.
FIGURE 6
FIGURE 6
Schematic illustration of the mechanisms underpinning skeletal muscle hypertrophy response to RET. After mechanical stimuli, phospholipase Cγ1 colocalizes around FAK and catalyzes the conversion of Ptdlns(4,5)P2 to PA that activates the HIPPO pathway (YAP and TAZ). YAP and TAZ promote the mTOR signaling pathway via the suppression of PTEN, regulate SC activation, and mediate the expression of L-type amino acid transporter 1 (LAT1). Interaction between Filamin-C and Bag3 increases the activation of the HIPPO pathway. Translational capacity and efficacy via ribosome are essential determinants for muscle hypertrophy. Besides increased gene expression during RE, the UTR of mRNA is regulated and closely correlated with muscle growth. The AR content is associated with RET-induced hypertrophy, whereas RE-induced acute changes in serum testosterone show no association with muscle growth. In human trials, RE-induced increased CK and inflammation factors have no effects on muscle hypertrophy. However, RET reduces resting concentration of inflammation cytokines in aging and clinical illness. AR, androgen receptor; ECM, extracellular matrix; FAK, focal adhesion kinase; IκBα, NF-κB inhibitor alpha; IKK, IκB kinase; LAT1, L-type amino acid transporter 1; mTOR, mechanistic target of rapamycin; MuRF 1, muscle RING-finger protein 1; NF-κB, nuclear factor kappa light-chain enhancer of activated B; PGE2, prostaglandin E2; PTEN, phosphatase and tensin homolog; RE, resistance exercise; RET, resistance exercise training; SC, satellite cells; TAZ, transcriptional coactivator with PDZ-binding motif; UTR, untranslated region; YAP, Yes-associated protein 1.

References

    1. Koeppel M, Mathis K, Schmitz KH, Wiskemann J. Muscle hypertrophy in cancer patients and survivors via strength training. A meta-analysis and meta-regression. Crit Rev Oncol Hematol. 2021;163:103371.
    1. McKendry J, Stokes T, McLeod JC, Phillips SM. Resistance exercise, aging, disuse, and muscle protein metabolism. Compr Physiol. 2021;11(3):2249–78.
    1. Glover EI Phillips SM Oates BR, et al. . Immobilization induces anabolic resistance in human myofibrillar protein synthesis with low and high dose amino acid infusion. J Physiol. 2008;586(24):6049–61.
    1. Cohen S, Nathan JA, Goldberg AL. Muscle wasting in disease: molecular mechanisms and promising therapies. Nat Rev Drug Discov. 2015;14(1):58–74.
    1. Wackerhage H, Schoenfeld BJ, Hamilton DL, Lehti M, Hulmi JJ. Stimuli and sensors that initiate skeletal muscle hypertrophy following resistance exercise. J Appl Physiol (1985). 2019;126(1):30–43.
    1. Gloerich M ten Klooster JP Vliem MJ, et al. . Rap2A links intestinal cell polarity to brush border formation. Nat Cell Biol. 2012;14(8):793–801.
    1. Meng Z Qiu Y Lin KC, et al. . RAP2 mediates mechanoresponses of the Hippo pathway. Nature. 2018;560(7720):655–60.
    1. Harvey K, Tapon N. The Salvador–Warts–Hippo pathway—an emerging tumour-suppressor network. Nat Rev Cancer. 2007;7(3):182–91.
    1. Pan D. Hippo signaling in organ size control. Genes Dev. 2007;21(8):886–97.
    1. Watt KI Judson R Medlow P, et al. . Yap is a novel regulator of C2C12 myogenesis. Biochem Biophys Res Commun. 2010;393(4):619–24.
    1. Jeong H Bae S An SY, et al. . TAZ as a novel enhancer of MyoD-mediated myogenic differentiation. FASEB J. 2010;24(9):3310–20.
    1. 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.
    1. Goodman CA, Dietz JM, Jacobs BL, McNally RM, You JS, Hornberger TA. Yes-associated protein is up-regulated by mechanical overload and is sufficient to induce skeletal muscle hypertrophy. FEBS Lett. 2015;589(13):1491–7.
    1. Hansen CG, Ng YL, Lam WL, Plouffe SW, Guan KL. The Hippo pathway effectors YAP and TAZ promote cell growth by modulating amino acid signaling to mTORC1. Cell Res. 2015;25(12):1299–313.
    1. Hornberger TA, Chu WK, Mak YW, Hsiung JW, Huang SA, Chien S. The role of phospholipase D and phosphatidic acid in the mechanical activation of mTOR signaling in skeletal muscle. Proc Natl Acad Sci U S A. 2006;103(12):4741–6.
    1. Furst DO, Osborn M, Nave R, Weber K. The organization of titin filaments in the half-sarcomere revealed by monoclonal antibodies in immunoelectron microscopy: a map of ten nonrepetitive epitopes starting at the Z line extends close to the M line. J Cell Biol. 1988;106(5):1563–72.
    1. Powers K, Schappacher-Tilp G, Jinha A, Leonard T, Nishikawa K, Herzog W. Titin force is enhanced in actively stretched skeletal muscle. J Exp Biol. 2014;217(Pt 20):3629–36.
    1. Puchner EM Alexandrovich A Kho AL, et al. . Mechanoenzymatics of titin kinase. Proc Natl Acad Sci U S A. 2008;105(36):13385–90.
    1. Kruger M, Kotter S. Titin, a central mediator for hypertrophic signaling, exercise-induced mechanosignaling and skeletal muscle remodeling. Front Physiol. 2016;7:76.
    1. Nishikawa K, Lindstedt SL, Hessel A, Mishra D. N2A Titin: signaling hub and mechanical switch in skeletal muscle. Int J Mol Sci. 2020;21(11):3974.
    1. Centner T Yano J Kimura E, et al. . Identification of muscle specific ring finger proteins as potential regulators of the titin kinase domain. J Mol Biol. 2001;306(4):717–26.
    1. Gregorio CC, Perry CN, McElhinny AS. Functional properties of the titin/connectin-associated proteins, the muscle-specific RING finger proteins (MURFs), in striated muscle. J Muscle Res Cell Motil. 2005;26(6–8):389–400.
    1. Ulbricht A Eppler FJ Tapia VE, et al. . Cellular mechanotransduction relies on tension-induced and chaperone-assisted autophagy. Curr Biol. 2013;23(5):430–5.
    1. Hoffman NJ Parker BL Chaudhuri R, et al. . Global phosphoproteomic analysis of human skeletal muscle reveals a network of exercise-regulated kinases and AMPK substrates. Cell Metab. 2015;22(5):922–35.
    1. Liu GY, Sabatini DM. mTOR at the nexus of nutrition, growth, ageing and disease. Nat Rev Mol Cell Biol. 2020;21(4):183–203.
    1. Szwed A, Kim E, Jacinto E. Regulation and metabolic functions of mTORC1 and mTORC2. Physiol Rev. 2021;101(3):1371–426.
    1. Reidy PT Borack MS Markofski MM, et al. . Post-absorptive muscle protein turnover affects resistance training hypertrophy. Eur J Appl Physiol. 2017;117(5):853–66.
    1. Mitchell CJ, Churchward-Venne TA, Bellamy L, Parise G, Baker SK, Phillips SM. Muscular and systemic correlates of resistance training–induced muscle hypertrophy. PLoS One. 2013;8(10):e78636.
    1. Mitchell CJ Churchward-Venne TA Parise G, et al. . Acute post-exercise myofibrillar protein synthesis is not correlated with resistance training–induced muscle hypertrophy in young men. PLoS One. 2014;9(2):e89431.
    1. Damas F Phillips SM Libardi CA, et al. . Resistance training–induced changes in integrated myofibrillar protein synthesis are related to hypertrophy only after attenuation of muscle damage. J Physiol. 2016;594(18):5209–22.
    1. Sancak Y, Bar-Peled L, Zoncu R, Markhard AL, Nada S, Sabatini DM. Ragulator–Rag complex targets mTORC1 to the lysosomal surface and is necessary for its activation by amino acids. Cell. 2010;141(2):290–303.
    1. Hodson N, Philp A. The importance of mTOR trafficking for human skeletal muscle translational control. Exerc Sport Sci Rev. 2019;47(1):46–53.
    1. Song Z Moore DR Hodson N, et al. . Resistance exercise initiates mechanistic target of rapamycin (mTOR) translocation and protein complex co-localisation in human skeletal muscle. Sci Rep. 2017;7(1):5028.
    1. Abou Sawan S van Vliet S Parel JT, et al. . Translocation and protein complex co-localization of mTOR is associated with postprandial myofibrillar protein synthesis at rest and after endurance exercise. Physiol Rep. 2018;6(5):e13628.
    1. Hannaian SJ Hodson N Abou Sawan S, et al. . Leucine-enriched amino acids maintain peripheral mTOR–Rheb localization independent of myofibrillar protein synthesis and mTORC1 signaling postexercise. J Appl Physiol (1985). 2020;129(1):133–43.
    1. Hodson N McGlory C Oikawa SY, et al. . Differential localization and anabolic responsiveness of mTOR complexes in human skeletal muscle in response to feeding and exercise. Am J Physiol Cell Physiol. 2017;313(6):C604–11.
    1. Abou Sawan S, Mazzulla M, Moore DR, Hodson N. More than just a garbage can: emerging roles of the lysosome as an anabolic organelle in skeletal muscle. Am J Physiol Cell Physiol. 2020;319(3):C561–8.
    1. Ogasawara R, Jensen TE, Goodman CA, Hornberger TA. Resistance exercise–induced hypertrophy: a potential role for rapamycin-insensitive mTOR. Exerc Sport Sci Rev. 2019;47(3):188–94.
    1. West DW Baehr LM Marcotte GR, et al. . Acute resistance exercise activates rapamycin-sensitive and -insensitive mechanisms that control translational activity and capacity in skeletal muscle. J Physiol. 2016;594(2):453–68.
    1. Ogasawara R Fujita S Hornberger TA, et al. . The role of mTOR signalling in the regulation of skeletal muscle mass in a rodent model of resistance exercise. Sci Rep. 2016;6:31142.
    1. Drummond MJ Fry CS Glynn EL, et al. . Rapamycin administration in humans blocks the contraction-induced increase in skeletal muscle protein synthesis. J Physiol. 2009;587(Pt 7):1535–46.
    1. Ogasawara R, Suginohara T. Rapamycin-insensitive mechanistic target of rapamycin regulates basal and resistance exercise–induced muscle protein synthesis. FASEB J. 2018;fj201701422R.
    1. Steinert ND Potts GK Wilson GM, et al. . Mapping of the contraction-induced phosphoproteome identifies TRIM28 as a significant regulator of skeletal muscle size and function. Cell Rep. 2021;34(9):108796.
    1. Henras AK, Plisson-Chastang C, O’Donohue MF, Chakraborty A, Gleizes PE. An overview of pre-ribosomal RNA processing in eukaryotes. Wiley Interdiscip Rev RNA. 2015;6(2):225–42.
    1. Mayer C, Grummt I. Ribosome biogenesis and cell growth: mTOR coordinates transcription by all three classes of nuclear RNA polymerases. Oncogene. 2006;25(48):6384–91.
    1. Chesley A, MacDougall JD, Tarnopolsky MA, Atkinson SA, Smith K. Changes in human muscle protein synthesis after resistance exercise. J Appl Physiol (1985). 1992;73(4):1383–8.
    1. Kirby TJ, Lee JD, England JH, Chaillou T, Esser KA, McCarthy JJ. Blunted hypertrophic response in age skeletal muscle is associated with decreased ribosome biogenesis. J Appl Physiol (1985). 2015;119(4):321–7.
    1. O’Neil D, Glowatz H, Schlumpberger M. Ribosomal RNA depletion for efficient use of RNA-seq capacity. Curr Protoc Mol Biol. 2013; Chapter 4:Unit 4.19.
    1. Goldberg AL, Etlinger JD, Goldspink DF, Jablecki C. Mechanism of work-induced hypertrophy of skeletal muscle. Med Sci Sports. 1975;7(3):185–98.
    1. Kotani T, Takegaki J, Tamura Y, Kouzaki K, Nakazato K, Ishii N. Repeated bouts of resistance exercise in rats alter mechanistic target of rapamycin complex 1 activity and ribosomal capacity but not muscle protein synthesis. Exp Physiol. 2021;106(9):1950–60.
    1. Hammarstrom D Ofsteng S Koll L, et al. . Benefits of higher resistance-training volume are related to ribosome biogenesis. J Physiol. 2020;598(3):543–65.
    1. Nader GA von Walden F Liu C, et al. . Resistance exercise training modulates acute gene expression during human skeletal muscle hypertrophy. J Appl Physiol (1985). 2014;116(6):693–702.
    1. 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(3):541–8.
    1. Brook MS Wilkinson DJ Mitchell WK, et al. . A novel D2O tracer method to quantify RNA turnover as a biomarker of de novo ribosomal biogenesis, in vitro, in animal models, and in human skeletal muscle. Am J Physiol Endocrinol Metab. 2017;313(6):E681–9.
    1. Brook MS Wilkinson DJ Mitchell WK, et al. . Synchronous deficits in cumulative muscle protein synthesis and ribosomal biogenesis underlie age-related anabolic resistance to exercise in humans. J Physiol. 2016;594(24):7399–417.
    1. Phillips BE Williams JP Gustafsson T, et al. . Molecular networks of human muscle adaptation to exercise and age. PLoS Genet. 2013;9(3):e1003389.
    1. Haun CT Vann CG Mobley CB, et al. . Pre-training skeletal muscle fiber size and predominant fiber type best predict hypertrophic responses to 6 weeks of resistance training in previously trained young men. Front Physiol. 2019;10:297.
    1. Mobley CB Haun CT Roberson PA, et al. . Biomarkers associated with low, moderate, and high vastus lateralis muscle hypertrophy following 12 weeks of resistance training. PLoS One. 2018;13(4):e0195203.
    1. Gibbons JG, Branco AT, Yu S, Lemos B. Ribosomal DNA copy number is coupled with gene expression variation and mitochondrial abundance in humans. Nat Commun. 2014;5:4850.
    1. Figueiredo VC Wen Y Alkner B, et al. . Genetic and epigenetic regulation of skeletal muscle ribosome biogenesis with exercise. J Physiol. 2021;599(13):3363–84.
    1. Atkinson G, Batterham AM. True and false interindividual differences in the physiological response to an intervention. Exp Physiol. 2015;100(6):577–88.
    1. Snapinn SM, Jiang Q. Responder analyses and the assessment of a clinically relevant treatment effect. Trials. 2007;8:31.
    1. Stokes T Timmons JA Crossland H, et al. . Molecular transducers of human skeletal muscle remodeling under different loading states. Cell Rep. 2020;32(5):107980.
    1. Robinson MM Dasari S Konopka AR, et al. . Enhanced protein translation underlies improved metabolic and physical adaptations to different exercise training modes in young and old humans. Cell Metab. 2017;25(3):581–92.
    1. Deveson IW, Hardwick SA, Mercer TR, Mattick JS. The dimensions, dynamics, and relevance of the mammalian noncoding transcriptome. Trends Genet. 2017;33(7):464–78.
    1. Jiang L Schlesinger F Davis CA, et al. . Synthetic spike-in standards for RNA-seq experiments. Genome Res. 2011;21(9):1543–51.
    1. Timmons JA, Szkop KJ, Gallagher IJ. Multiple sources of bias confound functional enrichment analysis of global -omics data. Genome Biol. 2015;16(1):186.
    1. Li JJ, Biggin MD. Gene expression. Statistics requantitates the central dogma. Science. 2015;347(6226):1066–7.
    1. Pillon NJ Gabriel BM Dollet L, et al. . Transcriptomic profiling of skeletal muscle adaptations to exercise and inactivity. Nat Commun. 2020;11(1):470.
    1. Raue U Trappe TA Estrem ST, et al. . Transcriptome signature of resistance exercise adaptations: mixed muscle and fiber type specific profiles in young and old adults. J Appl Physiol (1985). 2012;112(10):1625–36.
    1. MacInnis MJ, McGlory C, Gibala MJ, Phillips SM. Investigating human skeletal muscle physiology with unilateral exercise models: when one limb is more powerful than two. Appl Physiol Nutr Metab. 2017;42(6):563–70.
    1. Gharahdaghi N, Phillips BE, Szewczyk NJ, Smith K, Wilkinson DJ, Atherton PJ. Links between testosterone, oestrogen, and the growth hormone/insulin-like growth factor axis and resistance exercise muscle adaptations. Front Physiol. 2020;11:621226.
    1. Morton RW Sato K Gallaugher MPB, et al. . Muscle androgen receptor content but not systemic hormones is associated with resistance training–induced skeletal muscle hypertrophy in healthy, young men. Front Physiol. 2018;9:1373.
    1. West DW Kujbida GW Moore DR, et al. . Resistance exercise–induced increases in putative anabolic hormones do not enhance muscle protein synthesis or intracellular signalling in young men. J Physiol. 2009;587(Pt 21):5239–47.
    1. Morton RW Oikawa SY Wavell CG, et al. . Neither load nor systemic hormones determine resistance training–mediated hypertrophy or strength gains in resistance-trained young men. J Appl Physiol (1985). 2016;121(1):129–38.
    1. West DW Burd NA Tang JE, et al. . Elevations in ostensibly anabolic hormones with resistance exercise enhance neither training-induced muscle hypertrophy nor strength of the elbow flexors. J Appl Physiol (1985). 2010;108(1):60–7.
    1. Fryburg DA, Louard RJ, Gerow KE, Gelfand RA, Barrett EJ. Growth hormone stimulates skeletal muscle protein synthesis and antagonizes insulin’s antiproteolytic action in humans. Diabetes. 1992;41(4):424–9.
    1. West DW, Burd NA, Staples AW, Phillips SM. Human exercise-mediated skeletal muscle hypertrophy is an intrinsic process. Int J Biochem Cell Biol. 2010;42(9):1371–5.
    1. Yarasheski KE, Campbell JA, Smith K, Rennie MJ, Holloszy JO, Bier DM. Effect of growth hormone and resistance exercise on muscle growth in young men. Am J Physiol. 1992;262(3 Pt 1):E261–7.
    1. Friedlander AL Butterfield GE Moynihan S, et al. . One year of insulin-like growth factor I treatment does not affect bone density, body composition, or psychological measures in postmenopausal women. J Clin Endocrinol Metab. 2001;86(4):1496–503.
    1. Bhasin S Storer TW Berman N, et al. . The effects of supraphysiologic doses of testosterone on muscle size and strength in normal men. N Engl J Med. 1996;335(1):1–7.
    1. Young NR, Baker HW, Liu G, Seeman E. Body composition and muscle strength in healthy men receiving testosterone enanthate for contraception. J Clin Endocrinol Metab. 1993;77(4):1028–32.
    1. Bhasin S Storer TW Berman N, et al. . Testosterone replacement increases fat-free mass and muscle size in hypogonadal men. J Clin Endocrinol Metab. 1997;82(2):407–13.
    1. Katznelson L, Finkelstein JS, Schoenfeld DA, Rosenthal DI, Anderson EJ, Klibanski A. Increase in bone density and lean body mass during testosterone administration in men with acquired hypogonadism. J Clin Endocrinol Metab. 1996;81(12):4358–65.
    1. Gharahdaghi N Rudrappa S Brook MS, et al. . Testosterone therapy induces molecular programming augmenting physiological adaptations to resistance exercise in older men. J Cachexia Sarcopenia Muscle. 2019;10(6):1276–94.
    1. Schroeder ET, Villanueva M, West DD, Phillips SM. Are acute post-resistance exercise increases in testosterone, growth hormone, and IGF-1 necessary to stimulate skeletal muscle anabolism and hypertrophy? Med Sci Sports Exerc. 2013;45(11):2044–51.
    1. Willoughby DS, Taylor L. Effects of sequential bouts of resistance exercise on androgen receptor expression. Med Sci Sports Exerc. 2004;36(9):1499–506.
    1. Dobs AS, Meikle AW, Arver S, Sanders SW, Caramelli KE, Mazer NA. Pharmacokinetics, efficacy, and safety of a permeation-enhanced testosterone transdermal system in comparison with bi-weekly injections of testosterone enanthate for the treatment of hypogonadal men. J Clin Endocrinol Metab. 1999;84(10):3469–78.
    1. Cardaci TD, Machek SB, Wilburn DT, Heileson JL, Willoughby DS. High-load resistance exercise augments androgen receptor–DNA binding and Wnt/beta-catenin signaling without increases in serum/muscle androgens or androgen receptor content. Nutrients. 2020;12(12):3829.
    1. Roberts BM, Nuckols G, Krieger JW. Sex differences in resistance training: a systematic review and meta-analysis. J Strength Cond Res. 2020;34(5):1448–60.
    1. Hansen M. Female hormones: do they influence muscle and tendon protein metabolism? Proc Nutr Soc. 2018;77(1):32–41.
    1. Ahtiainen JP Hulmi JJ Kraemer WJ, et al. . Heavy resistance exercise training and skeletal muscle androgen receptor expression in younger and older men. Steroids. 2011;76(1–2):183–92.
    1. Peake JM, Neubauer O, Della Gatta PA, Nosaka K. Muscle damage and inflammation during recovery from exercise. J Appl Physiol (1985). 2017;122(3):559–70.
    1. Beaton LJ, Tarnopolsky MA, Phillips SM. Variability in estimating eccentric contraction-induced muscle damage and inflammation in humans. Can J Appl Physiol. 2002;27(5):516–26.
    1. Damas F Phillips SM Lixandrao ME, et al. . Early resistance training–induced increases in muscle cross-sectional area are concomitant with edema-induced muscle swelling. Eur J Appl Physiol. 2016;116(1):49–56.
    1. Snijders T Nederveen JP McKay BR, et al. . Satellite cells in human skeletal muscle plasticity. Front Physiol. 2015;6:283.
    1. Damas F Libardi CA Ugrinowitsch C, et al. . Early- and later-phases satellite cell responses and myonuclear content with resistance training in young men. PLoS One. 2018;13(1):e0191039.
    1. Moore DR, Phillips SM, Babraj JA, Smith K, Rennie MJ. 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(6):E1153–9.
    1. Roman W Pinheiro H Pimentel MR, et al. . Muscle repair after physiological damage relies on nuclear migration for cellular reconstruction. Science. 2021;374(6565):355–9.
    1. Alvarez AM, DeOcesano-Pereira C, Teixeira C, Moreira V. IL-1beta and TNF-alpha modulation of proliferated and committed myoblasts: IL-6 and COX-2-derived prostaglandins as key actors in the mechanisms involved. Cell. 2020;9(9):2005.
    1. Gao S, Durstine JL, Koh HJ, Carver WE, Frizzell N, Carson JA. Acute myotube protein synthesis regulation by IL-6-related cytokines. Am J Physiol Cell Physiol. 2017;313(5):C487–500.
    1. Krentz JR, Quest B, Farthing JP, Quest DW, Chilibeck PD. The effects of ibuprofen on muscle hypertrophy, strength, and soreness during resistance training. Appl Physiol Nutr Metab. 2008;33(3):470–5.
    1. Trappe TA Carroll CC Dickinson JM, et al. . Influence of acetaminophen and ibuprofen on skeletal muscle adaptations to resistance exercise in older adults. Am J Physiol Regul Integr Comp Physiol. 2011;300(3):R655–62.
    1. Schoenfeld BJ. Potential mechanisms for a role of metabolic stress in hypertrophic adaptations to resistance training. Sports Med. 2013;43(3):179–94.
    1. van der Knaap JA, Verrijzer CP. Undercover: gene control by metabolites and metabolic enzymes. Genes Dev. 2016;30(21):2345–69.
    1. MacDougall JD, Ray S, Sale DG, McCartney N, Lee P, Garner S. Muscle substrate utilization and lactate production. Can J Appl Physiol. 1999;24(3):209–15.
    1. Dankel SJ, Mattocks KT, Jessee MB, Buckner SL, Mouser JG, Loenneke JP. Do metabolites that are produced during resistance exercise enhance muscle hypertrophy? Eur J Appl Physiol. 2017;117(11):2125–35.
    1. Ampomah K Amano S Wages NP, et al. . Blood flow–restricted exercise does not induce a cross-transfer of effect: a randomized controlled trial. Med Sci Sports Exerc. 2019;51(9):1817–27.
    1. Bjornsen T Wernbom M Lovstad A, et al. . Delayed myonuclear addition, myofiber hypertrophy, and increases in strength with high-frequency low-load blood flow restricted training to volitional failure. J Appl Physiol (1985). 2019;126(3):578–92.
    1. Kefaloyianni E, Gaitanaki C, Beis I. ERK1/2 and p38-MAPK signalling pathways, through MSK1, are involved in NF-kappaB transactivation during oxidative stress in skeletal myoblasts. Cell Signal. 2006;18(12):2238–51.
    1. Morrison DK. MAP kinase pathways. Cold Spring Harb Perspect Biol. 2012;4(11):a011254.
    1. Fu X, Wang H, Hu P. Stem cell activation in skeletal muscle regeneration. Cell Mol Life Sci. 2015;72(9):1663–77.
    1. Hirvonen J, Nummela A, Rusko H, Rehunen S, Harkonen M. Fatigue and changes of ATP, creatine phosphate, and lactate during the 400-m sprint. Can J Sport Sci. 1992;17(2):141–4.
    1. Tesch PA, Colliander EB, Kaiser P. Muscle metabolism during intense, heavy-resistance exercise. Eur J Appl Physiol Occup Physiol. 1986;55(4):362–6.
    1. Lecker SH, Solomon V, Mitch WE, Goldberg AL. Muscle protein breakdown and the critical role of the ubiquitin–proteasome pathway in normal and disease states. J Nutr. 1999;129(1S Suppl):227S–37.
    1. Cheema BS, Chan D, Fahey P, Atlantis E. Effect of progressive resistance training on measures of skeletal muscle hypertrophy, muscular strength and health-related quality of life in patients with chronic kidney disease: a systematic review and meta-analysis. Sports Med. 2014;44(8):1125–38.
    1. Brienesse LA, Emerson MN. Effects of resistance training for people with Parkinson's disease: a systematic review. J Am Med Dir Assoc. 2013;14(4):236–41.
    1. Bland KA, Kouw IWK, van Loon LJC, Zopf EM, Fairman CM. Exercise-based interventions to counteract skeletal muscle mass loss in people with cancer: can we overcome the odds? Sports Med. 2022;52(5):1009–27.
    1. Pamoukdjian F, Bouillet T, Levy V, Soussan M, Zelek L, Paillaud E. Prevalence and predictive value of pre-therapeutic sarcopenia in cancer patients: a systematic review. Clin Nutr. 2018;37(4):1101–13.
    1. Bhatnagar S, Panguluri SK, Gupta SK, Dahiya S, Lundy RF, Kumar A. Tumor necrosis factor-alpha regulates distinct molecular pathways and gene networks in cultured skeletal muscle cells. PLoS One. 2010;5(10):e13262.
    1. Costamagna D, Costelli P, Sampaolesi M, Penna F. Role of inflammation in muscle homeostasis and myogenesis. Mediators Inflamm. 2015;2015:805172.
    1. Tuttle CSL, Thang LAN, Maier AB. Markers of inflammation and their association with muscle strength and mass: a systematic review and meta-analysis. Ageing Res Rev. 2020;64:101185.
    1. Hu Z, Wang H, Lee IH, Du J, Mitch WE. Endogenous glucocorticoids and impaired insulin signaling are both required to stimulate muscle wasting under pathophysiological conditions in mice. J Clin Invest. 2009;119(10):3059–69.
    1. Horber FF, Hoopeler H, Scheidegger JR, Grunig BE, Howald H, Frey FJ. Impact of physical training on the ultrastructure of midthigh muscle in normal subjects and in patients treated with glucocorticoids. J Clin Invest. 1987;79(4):1181–90.
    1. Lim C Dunford EC Valentino SE, et al. . Both traditional and stair climbing-based HIIT cardiac rehabilitation induce beneficial muscle adaptations. Med Sci Sports Exerc. 2021;53(6):1114–24.
    1. Mebarek S Komati H Naro F, et al. . Inhibition of de novo ceramide synthesis upregulates phospholipase D and enhances myogenic differentiation. J Cell Sci. 2007;120(Pt 3):407–16.
    1. Dube JJ Amati F Toledo FG, et al. . Effects of weight loss and exercise on insulin resistance, and intramyocellular triacylglycerol, diacylglycerol and ceramide. Diabetologia. 2011;54(5):1147–56.
    1. Schaap LA, Pluijm SM, Deeg DJ, Visser M. Inflammatory markers and loss of muscle mass (sarcopenia) and strength. Am J Med. 2006;119(6):526 e9–17.
    1. Jiao J, Demontis F. Skeletal muscle autophagy and its role in sarcopenia and organismal aging. Curr Opin Pharmacol. 2017;34:1–6.
    1. Verdijk LB, Snijders T, Drost M, Delhaas T, Kadi F, van Loon LJ. Satellite cells in human skeletal muscle; from birth to old age. Age (Dordr). 2014;36(2):545–7.
    1. Domingues-Faria C, Vasson MP, Goncalves-Mendes N, Boirie Y, Walrand S. Skeletal muscle regeneration and impact of aging and nutrition. Ageing Res Rev. 2016;26:22–36.
    1. Cade WT, Yarasheski KE. Metabolic and molecular aspects of sarcopenia. In: Runge MS, Patterson C, editors. Principles of Molecular Medicine. Totowa (NJ): Humana Press; 2006. pp. 529–34.
    1. McKay BR, Ogborn DI, Bellamy LM, Tarnopolsky MA, Parise G. Myostatin is associated with age-related human muscle stem cell dysfunction. FASEB J. 2012;26(6):2509–21.
    1. Wiedmer P Jung T Castro JP, et al. . Sarcopenia—molecular mechanisms and open questions. Ageing Res Rev. 2021;65:101200.
    1. Rom O, Reznick AZ. The role of E3 ubiquitin-ligases MuRF-1 and MAFbx in loss of skeletal muscle mass. Free Radic Biol Med. 2016;98:218–30.
    1. Tam SL, Gordon T. Mechanisms controlling axonal sprouting at the neuromuscular junction. J Neurocytol. 2003;32(5–8):961–74.
    1. Liberman K, Forti LN, Beyer I, Bautmans I. The effects of exercise on muscle strength, body composition, physical functioning and the inflammatory profile of older adults: a systematic review. Curr Opin Clin Nutr Metab Care. 2017;20(1):30–53.
    1. Grgic J, Garofolini A, Orazem J, Sabol F, Schoenfeld BJ, Pedisic Z. Effects of resistance training on muscle size and strength in very elderly adults: a systematic review and meta-analysis of randomized controlled trials. Sports Med. 2020;50(11):1983–99.
    1. Orchard P Manickam N Ventresca C, et al. . Human and rat skeletal muscle single-nuclei multi-omic integrative analyses nominate causal cell types, regulatory elements, and SNPs for complex traits. Genome Res. 2021;31(12):2258–75.
    1. Murach KA Dimet-Wiley AL Wen Y, et al. . Late-life exercise mitigates skeletal muscle epigenetic aging. Aging Cell. 2022;21(1):e13527.

Source: PubMed

3
订阅