Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

Control of satellite cell function in muscle regeneration and its disruption in ageing

Abstract

Skeletal muscle contains a designated population of adult stem cells, called satellite cells, which are generally quiescent. In homeostasis, satellite cells proliferate only sporadically and usually by asymmetric cell division to replace myofibres damaged by daily activity and maintain the stem cell pool. However, satellite cells can also be robustly activated upon tissue injury, after which they undergo symmetric divisions to generate new stem cells and numerous proliferating myoblasts that later differentiate to muscle cells (myocytes) to rebuild the muscle fibre, thereby supporting skeletal muscle regeneration. Recent discoveries show that satellite cells have a great degree of population heterogeneity, and that their cell fate choices during the regeneration process are dictated by both intrinsic and extrinsic mechanisms. Extrinsic cues come largely from communication with the numerous distinct stromal cell types in their niche, creating a dynamically interactive microenvironment. This Review discusses the role and regulation of satellite cells in skeletal muscle homeostasis and regeneration. In particular, we highlight the cell-intrinsic control of quiescence versus activation, the importance of satellite cell–niche communication, and deregulation of these mechanisms associated with ageing. The increasing understanding of how satellite cells are regulated will help to advance muscle regeneration and rejuvenation therapies.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Quiescent satellite cells, heterogeneity and niche composition.
Fig. 2: Overview of basic mechanisms controlling satellite cell quiescence and activation.
Fig. 3: Epigenetic, metabolic and proteostatic effectors regulating satellite cell fate.
Fig. 4: Decline of satellite cells in ageing.

Similar content being viewed by others

References

  1. Fuchs, E. & Blau, H. M. Tissue stem cells: architects of their niches. Cell Stem Cell 27, 532–556 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Mauro, A. Satellite cell of skeletal muscle fibers. J. Biophys. Biochem. Cytol. 9, 493–495 (1961).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Evano, B., Khalilian, S., Le Carrou, G., Almouzni, G. & Tajbakhsh, S. Dynamics of asymmetric and symmetric divisions of muscle stem cells in vivo and on artificial niches. Cell Rep. 30, 3195–3206.e7 (2020). This study shows that satellite cells divide symmetrically and asymmetrically in vivo, and that these cells can switch from asymmetric to symmetric cell division ex vivo.

    CAS  PubMed  Google Scholar 

  4. Robinson, D. C. L. et al. Negative elongation factor regulates muscle progenitor expansion for efficient myofiber repair and stem cell pool repopulation. Dev. Cell 56, 1014–1029.e7 (2021). This work demonstrates that proliferating satellite cells are required for stem cell replenishment on regenerated myofibres, with the negative elongation factor playing a major role.

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Pawlikowski, B., Dalla Betta, N., Antwine, T. & Olwin, B. B. Skeletal muscle stem cell self-renewal and differentiation kinetics revealed by EdU lineage tracing during regeneration. Preprint at biorXiv https://doi.org/10.1101/627851 (2019).

    Article  Google Scholar 

  6. Rodgers, J. T., Schroeder, M. D., Ma, C. & Rando, T. A. HGFA is an injury-regulated systemic factor that induces the transition of stem cells into GAlert. Cell Rep. 19, 479–486 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Rodgers, J. T. et al. mTORC1 controls the adaptive transition of quiescent stem cells from G0 to G(Alert). Nature 510, 393–396 (2014). This study shows two states of satellite cell quiescence (the G0 and GAlert states).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Takeda, S., Clemens, P. R. & Hoffman, E. P. Exon-skipping in Duchenne muscular dystrophy. J. Neuromuscul. Dis. https://doi.org/10.3233/JND-210682 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  9. Filippelli, R. L. & Chang, N. C. Empowering muscle stem cells for the treatment of duchenne muscular dystrophy. Cells Tissues Organs https://doi.org/10.1159/000514305 (2021).

    Article  PubMed  Google Scholar 

  10. Mackenzie, S. J., Nicolau, S., Connolly, A. M. & Mendell, J. R. Therapeutic approaches for Duchenne muscular dystrophy: old and new. Semin. Pediatr. Neurol. 37, 100877 (2021).

    PubMed  Google Scholar 

  11. Chemello, F., Bassel-Duby, R. & Olson, E. N. Correction of muscular dystrophies by CRISPR gene editing. J. Clin. Invest. 130, 2766–2776 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Hernandez-Hernandez, J. M., Garcia-Gonzalez, E. G., Brun, C. E. & Rudnicki, M. A. The myogenic regulatory factors, determinants of muscle development, cell identity and regeneration. Semin. Cell Dev. Biol. 72, 10–18 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Yin, H., Price, F. & Rudnicki, M. A. Satellite cells and the muscle stem cell niche. Physiol. Rev. 93, 23–67 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Beauchamp, J. R. et al. Expression of CD34 and Myf5 defines the majority of quiescent adult skeletal muscle satellite cells. J. Cell Biol. 151, 1221–1234 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Cornelison, D. D. & Wold, B. J. Single-cell analysis of regulatory gene expression in quiescent and activated mouse skeletal muscle satellite cells. Dev. Biol. 191, 270–283 (1997).

    CAS  PubMed  Google Scholar 

  16. Tierney, M. T. & Sacco, A. Satellite cell heterogeneity in skeletal muscle homeostasis. Trends Cell Biol. 26, 434–444 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Hughes, S. M. & Blau, H. M. Muscle fiber pattern is independent of cell lineage in postnatal rodent development. Cell 68, 659–671 (1992).

    CAS  PubMed  Google Scholar 

  18. DiMario, J. X., Fernyak, S. E. & Stockdale, F. E. Myoblasts transferred to the limbs of embryos are committed to specific fibre fates. Nature 362, 165–167 (1993).

    CAS  PubMed  Google Scholar 

  19. Sambasivan, R. et al. Distinct regulatory cascades govern extraocular and pharyngeal arch muscle progenitor cell fates. Dev. Cell 16, 810–821 (2009).

    CAS  PubMed  Google Scholar 

  20. Stuelsatz, P. et al. Extraocular muscle satellite cells are high performance myo-engines retaining efficient regenerative capacity in dystrophin deficiency. Dev. Biol. 397, 31–44 (2015).

    CAS  PubMed  Google Scholar 

  21. Evano, B. et al. Transcriptome and epigenome diversity and plasticity of muscle stem cells following transplantation. PLoS Genet. 16, e1009022 (2020). This study shows that adult mouse satellite cells from extraocular muscles and limb muscles have molecular and functional specificities. The identity of extraocular muscle satellite cells is driven mainly by the niche, as shown after engraftment in limb muscles, although a subpopulation expresses genes that resist the engraftment.

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Carrero-Rojas, G., Benitez-Temino, B., Pastor, A. M. & Davis-Lopez de Carrizosa, M. A. Muscle progenitors derived from extraocular muscles express higher levels of neurotrophins and their receptors than other cranial and limb muscles. Cells 9, 747 (2020).

    CAS  PubMed Central  Google Scholar 

  23. Kuang, S., Kuroda, K., Le Grand, F. & Rudnicki, M. A. Asymmetric self-renewal and commitment of satellite stem cells in muscle. Cell 129, 999–1010 (2007). This study shows that satellite cells are a heterogeneous population for Myf5 expression and satellite cells that have never expressed Myf5 are more naive and can reconstitute the satellite cell niche.

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Le Grand, F., Jones, A. E., Seale, V., Scime, A. & Rudnicki, M. A. Wnt7a activates the planar cell polarity pathway to drive the symmetric expansion of satellite stem cells. Cell Stem Cell 4, 535–547 (2009).

    PubMed  PubMed Central  Google Scholar 

  25. Rocheteau, P., Gayraud-Morel, B., Siegl-Cachedenier, I., Blasco, M. A. & Tajbakhsh, S. A subpopulation of adult skeletal muscle stem cells retains all template DNA strands after cell division. Cell 148, 112–125 (2012). Using Tg:Pax7-nGFP mice, this study shows that Pax7-nGFPhi satellite cells are more dormant than Pax7-nGFPlow satellite cells, and upon injury generate distinct daughter cells by asymmetrically segregating template DNA strands to the stem cell.

    CAS  PubMed  Google Scholar 

  26. Tierney, M. T., Stec, M. J., Rulands, S., Simons, B. D. & Sacco, A. Muscle stem cells exhibit distinct clonal dynamics in response to tissue repair and homeostatic aging. Cell Stem Cell 22, 119–127.e3 (2018).

    CAS  PubMed  Google Scholar 

  27. Chakkalakal, J. V., Jones, K. M., Basson, M. A. & Brack, A. S. The aged niche disrupts muscle stem cell quiescence. Nature 490, 355–360 (2012). This work demonstrates that the aged myofibre expresses FGF2, driving a subset of satellite cells to break quiescence and lose their self-renewal capacity.

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Chakkalakal, J. V. et al. Early forming label-retaining muscle stem cells require p27kip1 for maintenance of the primitive state. Development 141, 1649–1659 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Ono, Y. et al. Slow-dividing satellite cells retain long-term self-renewal ability in adult muscle. J. Cell Sci. 125, 1309–1317 (2012).

    CAS  PubMed  Google Scholar 

  30. Scaramozza, A. et al. Lineage tracing reveals a subset of reserve muscle stem cells capable of clonal expansion under stress. Cell Stem Cell 24, 944–957.e5 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Der Vartanian, A. et al. PAX3 confers functional heterogeneity in skeletal muscle stem cell responses to environmental stress. Cell Stem Cell 24, 958–973.e9 (2019).

    PubMed Central  Google Scholar 

  32. Liu, N. et al. A Twist2-dependent progenitor cell contributes to adult skeletal muscle. Nat. Cell Biol. 19, 202–213 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Chapman, M. R. et al. Sorting single satellite cells from individual myofibers reveals heterogeneity in cell-surface markers and myogenic capacity. Integr. Biol. 5, 692–702 (2013).

    CAS  Google Scholar 

  34. Garcia-Prat, L. et al. FoxO maintains a genuine muscle stem-cell quiescent state until geriatric age. Nat. Cell Biol. 22, 1307–1318 (2020). This study shows that CD34 serves to identify and isolate two distinct stem cell states (genuine and primed) with distinct functional fates. At geriatric age, the genuine state loses regenerative potential.

    CAS  PubMed  Google Scholar 

  35. Porpiglia, E. et al. High-resolution myogenic lineage mapping by single-cell mass cytometry. Nat. Cell Biol. 19, 558–567 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Dell’Orso, S. et al. Single cell analysis of adult mouse skeletal muscle stem cells in homeostatic and regenerative conditions. Development 146, dev174177 (2019).

    PubMed  PubMed Central  Google Scholar 

  37. Yartseva, V. et al. Heterogeneity of satellite cells implicates DELTA1/NOTCH2 signaling in self-renewal. Cell Rep. 30, 1491–1503.e6 (2020).

    CAS  PubMed  Google Scholar 

  38. Giordani, L. et al. High-dimensional single-cell cartography reveals novel skeletal muscle-resident cell populations. Mol. Cell 74, 609–621.e6 (2019). Using single-cell RNA-seq and mass cytometry, this study maps ten different cell types in adult mouse muscle and describes two new cell populations.

    CAS  PubMed  Google Scholar 

  39. De Micheli, A. J. et al. Single-cell analysis of the muscle stem cell hierarchy identifies heterotypic communication signals involved in skeletal muscle regeneration. Cell Rep. 30, 3583–3595.e5 (2020).

    PubMed  PubMed Central  Google Scholar 

  40. Barruet, E. et al. Functionally heterogeneous human satellite cells identified by single cell RNA sequencing. eLife 9, e51576 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Marg, A. et al. Human muscle-derived CLEC14A-positive cells regenerate muscle independent of PAX7. Nat. Commun. 10, 5776 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Harel, I. et al. Distinct origins and genetic programs of head muscle satellite cells. Dev. Cell 16, 822–832 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Ono, Y., Boldrin, L., Knopp, P., Morgan, J. E. & Zammit, P. S. Muscle satellite cells are a functionally heterogeneous population in both somite-derived and branchiomeric muscles. Dev. Biol. 337, 29–41 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Pietrosemoli, N. et al. Comparison of multiple transcriptomes exposes unified and divergent features of quiescent and activated skeletal muscle stem cells. Skelet. Muscle 7, 28 (2017).

    PubMed  PubMed Central  Google Scholar 

  45. Yue, L., Wan, R., Luan, S., Zeng, W. & Cheung, T. H. Dek modulates global intron retention during muscle stem cells quiescence exit. Dev. Cell 53, 661–676.e6 (2020).

    CAS  PubMed  Google Scholar 

  46. Baghdadi, M. B. et al. Reciprocal signalling by Notch-vollagen V-CALCR retains muscle stem cells in their niche. Nature 557, 714–718 (2018). This study shows that collagen V produced by satellite cells by binding to CALCR maintains them in a quiescent state.

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Montarras, D., L’Honore, A. & Buckingham, M. Lying low but ready for action: the quiescent muscle satellite cell. FEBS J. 280, 4036–4050 (2013).

    CAS  PubMed  Google Scholar 

  48. Cheung, T. H. et al. Maintenance of muscle stem-cell quiescence by microRNA-489. Nature 482, 524–528 (2012). This work identifies the miRNA pathway in general, and that of miR-489 in particular, in maintaining the quiescent state of satellite cells by suppressing DEK.

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Boutet, S. C. et al. Alternative polyadenylation mediates microRNA regulation of muscle stem cell function. Cell Stem Cell 10, 327–336 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Hausburg, M. A. et al. Post-transcriptional regulation of satellite cell quiescence by TTP-mediated mRNA decay. eLife 4, e03390 (2015).

    PubMed  PubMed Central  Google Scholar 

  51. de Morree, A. et al. Staufen1 inhibits MyoD translation to actively maintain muscle stem cell quiescence. Proc. Natl Acad. Sci. USA 114, E8996–E9005 (2017).

    PubMed  PubMed Central  Google Scholar 

  52. Crist, C. G., Montarras, D. & Buckingham, M. Muscle satellite cells are primed for myogenesis but maintain quiescence with sequestration of Myf5 mRNA targeted by microRNA-31 in mRNP granules. Cell Stem Cell 11, 118–126 (2012).

    CAS  PubMed  Google Scholar 

  53. Eliazer, S. et al. Wnt4 from the niche controls the mechano-properties and quiescent state of muscle stem cells. Cell Stem Cell 25, 654–665.e4 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Zhang, L. et al. The CalcR-PKA-Yap1 axis is critical for maintaining quiescence in muscle stem cells. Cell Rep. 29, 2154–2163.e5 (2019).

    CAS  PubMed  Google Scholar 

  55. Judson, R. N. et al. The Hippo pathway member Yap plays a key role in influencing fate decisions in muscle satellite cells. J. Cell Sci. 125, 6009–6019 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Zismanov, V. et al. Phosphorylation of eIF2alpha is a translational control mechanism regulating muscle stem cell quiescence and self-renewal. Cell Stem Cell 18, 79–90 (2016).

    CAS  PubMed  Google Scholar 

  57. Massenet, J., Gardner, E., Chazaud, B. & Dilworth, F. J. Epigenetic regulation of satellite cell fate during skeletal muscle regeneration. Skelet. Muscle 11, 4 (2021).

    PubMed  PubMed Central  Google Scholar 

  58. Liu, L. et al. Chromatin modifications as determinants of muscle stem cell quiescence and chronological aging. Cell Rep. 4, 189–204 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Sousa-Victor, P. et al. Geriatric muscle stem cells switch reversible quiescence into senescence. Nature 506, 316–321 (2014). This study reports that activation of the cell cycle inhibitor p16INK4A drives a fraction of satellite cells into a senescent state in geriatric mice in response to muscle injury.

    CAS  PubMed  Google Scholar 

  60. Juan, A. H. et al. Polycomb EZH2 controls self-renewal and safeguards the transcriptional identity of skeletal muscle stem cells. Genes Dev. 25, 789–794 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Garcia-Prat, L. et al. Autophagy maintains stemness by preventing senescence. Nature 529, 37–42 (2016). This work shows that autophagy is essential to maintain quiescence in satellite cells and that this activity declines with ageing, accounting for their regenerative decline.

    CAS  PubMed  Google Scholar 

  62. Boonsanay, V. et al. Regulation of skeletal muscle stem cell quiescence by Suv4-20h1-dependent facultative heterochromatin formation. Cell Stem Cell 18, 229–242 (2016).

    CAS  PubMed  Google Scholar 

  63. Luo, D. et al. Deltex2 represses MyoD expression and inhibits myogenic differentiation by acting as a negative regulator of Jmjd1c. Proc. Natl Acad. Sci. USA 114, E3071–E3080 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  64. McKinnell, I. W. et al. Pax7 activates myogenic genes by recruitment of a histone methyltransferase complex. Nat. Cell Biol. 10, 77–84 (2008).

    CAS  PubMed  Google Scholar 

  65. Zhang, K., Sha, J. & Harter, M. L. Activation of Cdc6 by MyoD is associated with the expansion of quiescent myogenic satellite cells. J. Cell Biol. 188, 39–48 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Ryall, J. G. et al. The NAD+-dependent SIRT1 deacetylase translates a metabolic switch into regulatory epigenetics in skeletal muscle stem cells. Cell Stem Cell 16, 171–183 (2015). This study shows that satellite cells undergo a metabolic switch from FAO to glycolysis during the transition from quiescence to proliferation, with is mediated by an NAD+–SIRT1 axis.

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Yucel, N. et al. Glucose metabolism drives histone acetylation landscape transitions that dictate muscle stem cell function. Cell Rep. 27, 3939–3955.e6 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Naito, M. et al. Dnmt3a regulates proliferation of muscle satellite cells via p57Kip2. PLoS Genet. 12, e1006167 (2016).

    PubMed  PubMed Central  Google Scholar 

  69. Dall’Agnese, A. et al. Transcription factor-directed re-wiring of chromatin architecture for somatic cell nuclear reprogramming toward trans-differentiation. Mol. Cell 76, 453–472.e8 (2019).

    PubMed  PubMed Central  Google Scholar 

  70. Zhang, N. et al. Muscle progenitor specification and myogenic differentiation are associated with changes in chromatin topology. Nat. Commun. 11, 6222 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Pala, F. et al. Distinct metabolic states govern skeletal muscle stem cell fates during prenatal and postnatal myogenesis. J. Cell Sci. 131, jcs212977 (2018). This study shows that quiescent satellite stem cells have perturbed OXPHOS during ageing and distinct requirements for peroxisomal and mitochondrial FAO at different myogenic stages.

    PubMed  PubMed Central  Google Scholar 

  72. Cerletti, M., Jang, Y. C., Finley, L. W., Haigis, M. C. & Wagers, A. J. Short-term calorie restriction enhances skeletal muscle stem cell function. Cell Stem Cell 10, 515–519 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Garcia-Prat, L., Sousa-Victor, P. & Munoz-Canoves, P. Proteostatic and metabolic control of stemness. Cell stem Cell 20, 593–608 (2017).

    CAS  PubMed  Google Scholar 

  74. Chen, F. et al. YY1 regulates skeletal muscle regeneration through controlling metabolic reprogramming of satellite cells. EMBO J. 8, e99727 (2019).

    Google Scholar 

  75. Pallafacchina, G. et al. An adult tissue-specific stem cell in its niche: a gene profiling analysis of in vivo quiescent and activated muscle satellite cells. Stem Cell Res. 4, 77–91 (2010).

    CAS  PubMed  Google Scholar 

  76. Li, H. et al. Muscle-secreted granulocyte colony-stimulating factor functions as metabolic niche factor ameliorating loss of muscle stem cells in aged mice. EMBO J. 38, e102154 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  77. Wang, G. et al. p110α of PI3K is necessary and sufficient for quiescence exit in adult muscle satellite cells. EMBO J. 37, e98239 (2018).

    PubMed  PubMed Central  Google Scholar 

  78. Wu, Z. et al. p38 and extracellular signal-regulated kinases regulate the myogenic program at multiple steps. Mol. Cell Biol. 20, 3951–3964 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  79. Yue, F. et al. Conditional loss of Pten in myogenic progenitors leads to postnatal skeletal muscle hypertrophy but age-dependent exhaustion of satellite cells. Cell Rep. 17, 2340–2353 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  80. Pavlidou, T. et al. Metformin delays satellite cell activation and maintains quiescence. Stem Cell Int. 2019, 5980465 (2019).

    Google Scholar 

  81. Tang, A. H. & Rando, T. A. Induction of autophagy supports the bioenergetic demands of quiescent muscle stem cell activation. EMBO J. 33, 2782–2797 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  82. Bustos, F. et al. NEDD4 regulates PAX7 levels promoting activation of the differentiation program in skeletal muscle precursors. Stem Cell 33, 3138–3151 (2015).

    CAS  Google Scholar 

  83. Kitajima, Y. et al. The ubiquitin-proteasome system is indispensable for the maintenance of muscle stem cells. Stem Cell Rep. 11, 1523–1538 (2018).

    CAS  Google Scholar 

  84. Fiacco, E. et al. Autophagy regulates satellite cell ability to regenerate normal and dystrophic muscles. Cell Death Differ. 23, 1839–1849 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  85. Lahiri, V., Hawkins, W. D. & Klionsky, D. J. Watch what you (self-) eat: autophagic mechanisms that modulate metabolism. Cell Metab. 29, 803–826 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  86. Rozo, M., Li, L. & Fan, C. M. Targeting beta1-integrin signaling enhances regeneration in aged and dystrophic muscle in mice. Nat. Med. 22, 889–896 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  87. Webster, M. T., Manor, U., Lippincott-Schwartz, J. & Fan, C. M. Intravital imaging reveals ghost fibers as architectural units guiding myogenic progenitors during regeneration. Cell Stem Cell 18, 243–252 (2016).

    CAS  PubMed  Google Scholar 

  88. Gurevich, D. B. et al. Asymmetric division of clonal muscle stem cells coordinates muscle regeneration in vivo. Science 353, aad9969 (2016).

    PubMed  Google Scholar 

  89. Bentzinger, C. F. et al. Fibronectin regulates Wnt7a signaling and satellite cell expansion. Cell Stem Cell 12, 75–87 (2013). This work shows that activated satellite cells remodel their niche through autologous expression of fibronectin to stimulate WNT7A signalling through the FZD7–syndecan 4 co-receptor complex.

    CAS  PubMed  PubMed Central  Google Scholar 

  90. Dumont, N. A. et al. Dystrophin expression in muscle stem cells regulates their polarity and asymmetric division. Nat. Med. 21, 1455–1463 (2015). This work shows that dystrophin in satellite cells promotes asymmetric cell division and therefore cell proliferation.

    CAS  PubMed  PubMed Central  Google Scholar 

  91. Chang, N. C. et al. The dystrophin glycoprotein complex regulates the epigenetic activation of muscle stem cell commitment. Cell Stem Cell 22, 755–768.e6 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  92. Yennek, S., Burute, M., Thery, M. & Tajbakhsh, S. Cell adhesion geometry regulates non-random DNA segregation and asymmetric cell fates in mouse skeletal muscle stem cells. Cell Rep. 7, 961–970 (2014).

    CAS  PubMed  Google Scholar 

  93. Bernet, J. D. et al. p38 MAPK signaling underlies a cell-autonomous loss of stem cell self-renewal in skeletal muscle of aged mice. Nat. Med. 20, 265–271 (2014). This work shows that old satellite cells fail to self-renew by symmetrically distributing p38 MAPK, generating two daughter committed progenitors for myogenic differentiation.

    CAS  PubMed  PubMed Central  Google Scholar 

  94. Chen, Z. et al. Forkhead xox M1 transcriptionally regulates the expression of long noncoding RNAs Snhg8 and Gm26917 to promote proliferation and survival of muscle satellite cells. Stem Cell 36, 1097–1108 (2018).

    CAS  Google Scholar 

  95. Diao, Y. et al. Pax3/7BP is a Pax7- and Pax3-binding protein that regulates the proliferation of muscle precursor cells by an epigenetic mechanism. Cell stem Cell 11, 231–241 (2012).

    CAS  PubMed  Google Scholar 

  96. Sreenivasan, K. et al. Attenuated epigenetic suppression of muscle stem cell necroptosis is required for efficient regeneration of dystrophic muscles. Cell Rep. 31, 107652 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  97. Sreenivasan, K. et al. CHD4 ensures stem cell lineage fidelity during skeletal muscle regeneration. Stem Cell Reports 16, 2089–2098 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  98. Chen, X. et al. Translational control by DHX36 binding to 5’UTR G-quadruplex is essential for muscle stem-cell regenerative functions. Nat. Commun. 12, 5043 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  99. Choi, S., Ferrari, G. & Tedesco, F. S. Cellular dynamics of myogenic cell migration: molecular mechanisms and implications for skeletal muscle cell therapies. EMBO Mol. Med. 12, e12357 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  100. Dumont, N. A., Wang, Y. X. & Rudnicki, M. A. Intrinsic and extrinsic mechanisms regulating satellite cell function. Development 142, 1572–1581 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  101. Perdiguero, E. et al. Genetic analysis of p38 MAP kinases in myogenesis: fundamental role of p38alpha in abrogating myoblast proliferation. EMBO J. 26, 1245–1256 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  102. Zetser, A., Gredinger, E. & Bengal, E. p38 mitogen-activated protein kinase pathway promotes skeletal muscle differentiation. Participation of the Mef2c transcription factor. J. Biol. Chem. 274, 5193–5200 (1999).

    CAS  PubMed  Google Scholar 

  103. Liu, Q. C. et al. Comparative expression profiling identifies differential roles for myogenin and p38alpha MAPK signaling in myogenesis. J. Mol. Cell Biol. 4, 386–397 (2012). This work shows that QSCs and ASCs have distinct epigenetic landscapes.

    CAS  PubMed  PubMed Central  Google Scholar 

  104. Yu, X. et al. Long non-coding RNA Linc-RAM enhances myogenic differentiation by interacting with MyoD. Nat. Commun. 8, 14016 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  105. Das, S. et al. ATP citrate lyase regulates myofiber differentiation and increases regeneration by altering histone acetylation. Cell Rep. 21, 3003–3011 (2017).

    CAS  PubMed  Google Scholar 

  106. Chen, J. F. et al. The role of microRNA-1 and microRNA-133 in skeletal muscle proliferation and differentiation. Nat. Genet. 38, 228–233 (2006).

    CAS  PubMed  Google Scholar 

  107. Kim, H. K., Lee, Y. S., Sivaprasad, U., Malhotra, A. & Dutta, A. Muscle-specific microRNA miR-206 promotes muscle differentiation. J. Cell Biol. 174, 677–687 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  108. Dey, B. K., Gagan, J. & Dutta, A. miR-206 and -486 induce myoblast differentiation by downregulating Pax7. Mol. Cell Biol. 31, 203–214 (2011).

    CAS  PubMed  Google Scholar 

  109. Peng, X. L. et al. MyoD- and FoxO3-mediated hotspot interaction orchestrates super-enhancer activity during myogenic differentiation. Nucleic Acids Res. 45, 8785–8805 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  110. Olguin, H. C., Yang, Z., Tapscott, S. J. & Olwin, B. B. Reciprocal inhibition between Pax7 and muscle regulatory factors modulates myogenic cell fate determination. J. Cell Biol. 177, 769–779 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  111. Dong, A. et al. A long noncoding RNA, LncMyoD, modulates chromatin accessibility to regulate muscle stem cell myogenic lineage progression. Proc. Natl Acad. Sci. USA 117, 32464–32475 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  112. Sartorelli, V. & Puri, P. L. Shaping gene expression by landscaping chromatin architecture: lessons from a master. Mol. Cell 71, 375–388 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  113. Singh, K. & Dilworth, F. J. Differential modulation of cell cycle progression distinguishes members of the myogenic regulatory factor family of transcription factors. FEBS J. 280, 3991–4003 (2013).

    CAS  PubMed  Google Scholar 

  114. Bi, P. et al. Control of muscle formation by the fusogenic micropeptide myomixer. Science 356, 323–327 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  115. Hori, S., Hiramuki, Y., Nishimura, D., Sato, F. & Sehara-Fujisawa, A. PDH-mediated metabolic flow is critical for skeletal muscle stem cell differentiation and myotube formation during regeneration in mice. FASEB J. 33, 8094–8109 (2019).

    CAS  PubMed  Google Scholar 

  116. Lyons, C. N., Leary, S. C. & Moyes, C. D. Bioenergetic remodeling during cellular differentiation: changes in cytochrome c oxidase regulation do not affect the metabolic phenotype. Biochem. Cell Biol. 82, 391–399 (2004).

    CAS  PubMed  Google Scholar 

  117. Moyes, C. D. Using humans to study the physiological evolution of energy metabolism: a tribute to Peter Hochachka. Comp. Biochem. Physiol. B Biochem Mol. Biol. 139, 487–494 (2004).

    PubMed  Google Scholar 

  118. Hoffmann, C. et al. The effect of differentiation and TGFbeta on mitochondrial respiration and mitochondrial enzyme abundance in cultured primary human skeletal muscle cells. Sci. Rep. 8, 737 (2018).

    PubMed  PubMed Central  Google Scholar 

  119. Sin, J. et al. Mitophagy is required for mitochondrial biogenesis and myogenic differentiation of C2C12 myoblasts. Autophagy 12, 369–380 (2016).

    CAS  PubMed  Google Scholar 

  120. Shintaku, J. et al. MyoD regulates skeletal muscle oxidative metabolism cooperatively with alternative NF-kappaB. Cell Rep. 17, 514–526 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  121. Millay, D. P. et al. Myomaker is a membrane activator of myoblast fusion and muscle formation. Nature 499, 301–305 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  122. Quinn, M. E. et al. Myomerger induces fusion of non-fusogenic cells and is required for skeletal muscle development. Nat. Commun. 8, 15665 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  123. Zhang, Q. et al. The microprotein Minion controls cell fusion and muscle formation. Nat. Commun. 8, 15664 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  124. Cosgrove, B. D. et al. Rejuvenation of the muscle stem cell population restores strength to injured aged muscles. Nat. Med. 20, 255–264 (2014). This work demonstrates that p38 MAPK inhibition, together with cell culture on soft hydrogels, enhances the regenerative functions of old satellite cells.

    CAS  PubMed  PubMed Central  Google Scholar 

  125. Price, F. D. et al. Inhibition of JAK-STAT signaling stimulates adult satellite cell function. Nat. Med. 20, 1174–1181 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  126. Tierney, M. T. et al. STAT3 signaling controls satellite cell expansion and skeletal muscle repair. Nat. Med. 20, 1182–1186 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  127. Shea, K. L. et al. Sprouty1 regulates reversible quiescence of a self-renewing adult muscle stem cell pool during regeneration. Cell Stem Cell 6, 117–129 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  128. Gopinath, S. D., Webb, A. E., Brunet, A. & Rando, T. A. FOXO3 promotes quiescence in adult muscle stem cells during the process of self-renewal. Stem Cell Rep. 2, 414–426 (2014).

    CAS  Google Scholar 

  129. Brohl, D. et al. Colonization of the satellite cell niche by skeletal muscle progenitor cells depends on Notch signals. Dev. Cell 23, 469–481 (2012).

    PubMed  Google Scholar 

  130. Wen, Y. et al. Constitutive Notch activation upregulates Pax7 and promotes the self-renewal of skeletal muscle satellite cells. Mol. Cell Biol. 32, 2300–2311 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  131. Troy, A. et al. Coordination of satellite cell activation and self-renewal by Par-complex-dependent asymmetric activation of p38alpha/beta MAPK. Cell Stem Cell 11, 541–553 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  132. Wang, Y. X. et al. EGFR-Aurka signaling rescues polarity and regeneration defects in dystrophin-deficient muscle stem cells by increasing asymmetric divisions. Cell Stem Cell 24, 419–432.e6 (2019).

    PubMed  PubMed Central  Google Scholar 

  133. Goel, A. J., Rieder, M. K., Arnold, H. H., Radice, G. L. & Krauss, R. S. Niche cadherins control the quiescence-to-activation transition in muscle stem cells. Cell Rep. 21, 2236–2250 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  134. Sampath, S. C. et al. Induction of muscle stem cell quiescence by the secreted niche factor oncostatin M. Nat. Commun. 9, 1531 (2018).

    PubMed  PubMed Central  Google Scholar 

  135. Rayagiri, S. S. et al. Basal lamina remodeling at the skeletal muscle stem cell niche mediates stem cell self-renewal. Nat. Commun. 9, 1075 (2018).

    PubMed  PubMed Central  Google Scholar 

  136. Tierney, M. T. et al. Autonomous extracellular matrix remodeling controls a progressive adaptation in muscle stem cell regenerative capacity during development. Cell Rep. 14, 1940–1952 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  137. Urciuolo, A. et al. Collagen VI regulates satellite cell self-renewal and muscle regeneration. Nat. Commun. 4, 1964 (2013).

    PubMed  Google Scholar 

  138. Ratnayake, D. et al. Macrophages provide a transient muscle stem cell niche via NAMPT secretion. Nature 591, 281–287 (2021). This study reveals that a specific subset of macrophages ‘dwell’ within the muscle injury, establishing a transient but necessary niche for satellite cell proliferation.

    CAS  PubMed  Google Scholar 

  139. Lukjanenko, L. et al. Loss of fibronectin from the aged stem cell niche affects the regenerative capacity of skeletal muscle in mice. Nat. Med. 22, 897–905 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  140. Liu, W. & Chakkalakal, J. V. The composition, development, and regeneration of neuromuscular junctions. Curr. Top. Dev. Biol. 126, 99–124 (2018).

    PubMed  Google Scholar 

  141. Madaro, L. et al. Denervation-activated STAT3-IL-6 signalling in fibro-adipogenic progenitors promotes myofibres atrophy and fibrosis. Nat. Cell Biol. 20, 917–927 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  142. Liu, W., Wei-LaPierre, L., Klose, A., Dirksen, R. T. & Chakkalakal, J. V. Inducible depletion of adult skeletal muscle stem cells impairs the regeneration of neuromuscular junctions. eLife 4, e09221 (2015).

    PubMed Central  Google Scholar 

  143. Joe, A. W. et al. Muscle injury activates resident fibro/adipogenic progenitors that facilitate myogenesis. Nat. Cell Biol. 12, 153–163 (2010). This work reports a new subpopulation of FAPs resident in skeletal muscle tissue.

    CAS  PubMed  PubMed Central  Google Scholar 

  144. Uezumi, A., Fukada, S., Yamamoto, N., Takeda, S. & Tsuchida, K. Mesenchymal progenitors distinct from satellite cells contribute to ectopic fat cell formation in skeletal muscle. Nat. Cell Biol. 12, 143–152 (2010). This work identifies PDGFRα+ mesenchymal progenitors in skeletal muscle with adipogenic differentiation potential.

    CAS  PubMed  Google Scholar 

  145. Wosczyna, M. N., Biswas, A. A., Cogswell, C. A. & Goldhamer, D. J. Multipotent progenitors resident in the skeletal muscle interstitium exhibit robust BMP-dependent osteogenic activity and mediate heterotopic ossification. J. Bone Miner. Res. 27, 1004–1017 (2012).

    CAS  PubMed  Google Scholar 

  146. Wosczyna, M. N. et al. Mesenchymal stromal cells are required for regeneration and homeostatic maintenance of skeletal muscle. Cell Rep. 27, 2029–2035.e5 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  147. Shang, M. et al. Macrophage-derived glutamine boosts satellite cells and muscle regeneration. Nature 587, 626–631 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  148. Uezumi, A. et al. Fibrosis and adipogenesis originate from a common mesenchymal progenitor in skeletal muscle. J. Cell Sci. 124, 3654–3664 (2011).

    CAS  PubMed  Google Scholar 

  149. Kopinke, D., Roberson, E. C. & Reiter, J. F. Ciliary hedgehog signaling restricts injury-induced adipogenesis. Cell 170, 340–351.e2 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  150. Wosczyna, M. N. et al. Targeting microRNA-mediated gene repression limits adipogenic conversion of skeletal muscle mesenchymal stromal cells. Cell Stem Cell 28, 1323–1334.e8 (2021).

    CAS  PubMed  Google Scholar 

  151. Malecova, B. et al. Dynamics of cellular states of fibro-adipogenic progenitors during myogenesis and muscular dystrophy. Nat. Commun. 9, 3670 (2018).

    PubMed  PubMed Central  Google Scholar 

  152. Lemos, D. R. et al. Nilotinib reduces muscle fibrosis in chronic muscle injury by promoting TNF-mediated apoptosis of fibro/adipogenic progenitors. Nat. Med. 21, 786–794 (2015). This study demonstrates that sequential interactions between multipotent FAPs and infiltrating macrophages determine the outcome of the muscle regeneration process.

    CAS  PubMed  Google Scholar 

  153. Vidal, B. et al. Fibrinogen drives dystrophic muscle fibrosis via a TGFbeta/alternative macrophage activation pathway. Genes Dev. 22, 1747–1752 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  154. Tidball, J. G. Regulation of muscle growth and regeneration by the immune system. Nat. Rev. Immunol. 17, 165–178 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  155. Chazaud, B. Inflammation and skeletal muscle regeneration: leave it to the macrophages! Trends Immunol. 41, 481–492 (2020).

    CAS  PubMed  Google Scholar 

  156. Arnold, L. et al. Inflammatory monocytes recruited after skeletal muscle injury switch into antiinflammatory macrophages to support myogenesis. J. Exp. Med. 204, 1057–1069 (2007). This work shows that injured skeletal muscle recruits inflammatory monocytes that subsequently convert to anti-inflammatory macrophages that stimulate myogenic differentiation and muscle growth.

    CAS  PubMed  PubMed Central  Google Scholar 

  157. Wang, H. et al. Altered macrophage phenotype transition impairs skeletal muscle regeneration. Am. J. Pathol. 184, 1167–1184 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  158. Perdiguero, E. et al. p38/MKP-1-regulated AKT coordinates macrophage transitions and resolution of inflammation during tissue repair. J. Cell Biol. 195, 307–322 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  159. Heredia, J. E. et al. Type 2 innate signals stimulate fibro/adipogenic progenitors to facilitate muscle regeneration. Cell 153, 376–388 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  160. Baht, G. S. et al. Meteorin-like facilitates skeletal muscle repair through a Stat3/IGF-1 mechanism. Nat. Metab. 2, 278–289 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  161. Burzyn, D. et al. A special population of regulatory T cells potentiates muscle repair. Cell 155, 1282–1295 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  162. Villalta, S. A. et al. Regulatory T cells suppress muscle inflammation and injury in muscular dystrophy. Sci. Transl Med. 6, 258ra142 (2014).

    PubMed  PubMed Central  Google Scholar 

  163. Christov, C. et al. Muscle satellite cells and endothelial cells: close neighbors and privileged partners. Mol. Biol. Cell 18, 1397–1409 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  164. Roberts, P. & McGeachie, J. K. Endothelial cell activation during angiogenesis in freely transplanted skeletal muscles in mice and its relationship to the onset of myogenesis. J. Anat. 169, 197–207 (1990).

    CAS  PubMed  PubMed Central  Google Scholar 

  165. Bryan, B. A. et al. Coordinated vascular endothelial growth factor expression and signaling during skeletal myogenic differentiation. Mol. Biol. Cell 19, 994–1006 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  166. Tatsumi, R., Anderson, J. E., Nevoret, C. J., Halevy, O. & Allen, R. E. HGF/SF is present in normal adult skeletal muscle and is capable of activating satellite cells. Dev. Biol. 194, 114–128 (1998).

    CAS  PubMed  Google Scholar 

  167. Latroche, C. et al. Coupling between myogenesis and angiogenesis during skeletal muscle regeneration is stimulated by restorative macrophages. Stem Cell Rep. 9, 2018–2033 (2017).

    CAS  Google Scholar 

  168. Verma, M. et al. Muscle satellite cell cross-talk with a vascular niche maintains quiescence via VEGF and notch signaling. Cell Stem Cell 23, 530–543.e9 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  169. Shefer, G., Rauner, G., Yablonka-Reuveni, Z. & Benayahu, D. Reduced satellite cell numbers and myogenic capacity in aging can be alleviated by endurance exercise. PLoS ONE 5, e13307 (2010).

    PubMed  PubMed Central  Google Scholar 

  170. Shefer, G., Van de Mark, D. P., Richardson, J. B. & Yablonka-Reuveni, Z. Satellite-cell pool size does matter: defining the myogenic potency of aging skeletal muscle. Dev. Biol. 294, 50–66 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  171. Brack, A. S., Bildsoe, H. & Hughes, S. M. Evidence that satellite cell decrement contributes to preferential decline in nuclear number from large fibres during murine age-related muscle atrophy. J. Cell Sci. 118, 4813–4821 (2005).

    CAS  PubMed  Google Scholar 

  172. Haller, S. et al. mTORC1 activation during repeated regeneration impairs somatic stem cell maintenance. Cell Stem Cell 21, 806–818.e5 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  173. Sahu, A. et al. Age-related declines in alpha-Klotho drive progenitor cell mitochondrial dysfunction and impaired muscle regeneration. Nat. Commun. 9, 4859 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  174. Baraibar, M. A. et al. Impaired energy metabolism of senescent muscle satellite cells is associated with oxidative modifications of glycolytic enzymes. Aging 8, 3375–3389 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  175. Sousa-Victor, P., Perdiguero, E. & Munoz-Canoves, P. Geroconversion of aged muscle stem cells under regenerative pressure. Cell Cycle 13, 3183–3190 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  176. Childs, B. G., Baker, D. J., Kirkland, J. L., Campisi, J. & van Deursen, J. M. Senescence and apoptosis: dueling or complementary cell fates? EMBO Rep. 15, 1139–1153 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  177. Coppe, J. P. et al. Senescence-associated secretory phenotypes reveal cell-nonautonomous functions of oncogenic RAS and the p53 tumor suppressor. PLoS Biol. 6, 2853–2868 (2008).

    CAS  PubMed  Google Scholar 

  178. Franceschi, C. & Campisi, J. Chronic inflammation (inflammaging) and its potential contribution to age-associated diseases. J. Gerontol. A 69 (Suppl. 1), 4–9 (2014).

    Google Scholar 

  179. White, J. P. et al. The AMPK/p27(Kip1) axis regulates autophagy/apoptosis decisions in aged skeletal muscle stem cells. Stem Cell Rep. 11, 425–439 (2018).

    CAS  Google Scholar 

  180. Fry, C. S. et al. Inducible depletion of satellite cells in adult, sedentary mice impairs muscle regenerative capacity without affecting sarcopenia. Nat. Med. 21, 76–80 (2015).

    CAS  PubMed  Google Scholar 

  181. Englund, D. A. et al. Depletion of resident muscle stem cells negatively impacts running volume, physical function, and muscle fiber hypertrophy in response to lifelong physical activity. Am. J. Physiol. Cell Physiol. 318, C1178–C1188 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  182. Liu, W. et al. Loss of adult skeletal muscle stem cells drives age-related neuromuscular junction degeneration. eLife 6, e26464 (2017).

    PubMed  PubMed Central  Google Scholar 

  183. Conboy, I. M., Conboy, M. J., Smythe, G. M. & Rando, T. A. Notch-mediated restoration of regenerative potential to aged muscle. Science 302, 1575–1577 (2003).

    CAS  Google Scholar 

  184. Conboy, I. M. et al. Rejuvenation of aged progenitor cells by exposure to a young systemic environment. Nature 433, 760–764 (2005).

    CAS  Google Scholar 

  185. Zhang, H. et al. NAD+ repletion improves mitochondrial and stem cell function and enhances life span in mice. Science 352, 1436–1443 (2016).

    CAS  PubMed  Google Scholar 

  186. Solanas, G. et al. Aged stem cells reprogram their daily rhythmic functions to adapt to stress. Cell 170, 678–692.e20 (2017).

    CAS  PubMed  Google Scholar 

  187. Carlson, M. E., Hsu, M. & Conboy, I. M. Imbalance between pSmad3 and Notch induces CDK inhibitors in old muscle stem cells. Nature 454, 528–532 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  188. Brett, J. O. et al. Exercise rejuvenates quiescent skeletal muscle stem cells in old mice through restoration of cyclin D1. Nat. Metab. 2, 307–317 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  189. Liu, L. et al. Impaired notch signaling leads to a decrease in p53 activity and mitotic catastrophe in aged muscle stem cells. Cell Stem Cell 23, 544–556.e4 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  190. Brack, A. S. et al. Increased Wnt signaling during aging alters muscle stem cell fate and increases fibrosis. Science 317, 807–810 (2007).

    CAS  Google Scholar 

  191. An, Y. et al. A molecular switch regulating cell fate choice between muscle progenitor cells and brown adipocytes. Dev. Cell 41, 382–391.e5 (2017).

    CAS  PubMed  Google Scholar 

  192. Schworer, S. et al. Epigenetic stress responses induce muscle stem-cell ageing by Hoxa9 developmental signals. Nature 540, 428–432 (2016). This study shows that the epigenetic stress response in satellite cells differs between young and old mice and identifies HOXA9 as a central factor that activates several developmental pathways in aged proliferating satellite cells.

    PubMed  Google Scholar 

  193. Hernando-Herraez, I. et al. Ageing affects DNA methylation drift and transcriptional cell-to-cell variability in mouse muscle stem cells. Nat. Commun. 10, 4361 (2019).

    PubMed  PubMed Central  Google Scholar 

  194. Gutmann, E. & Carlson, B. M. Regeneration and transplantation of muscles in old rats and between young and old rats. Life Sci. 18, 109–114 (1976).

    CAS  PubMed  Google Scholar 

  195. Carlson, B. M. & Faulkner, J. A. Muscle transplantation between young and old rats: age of host determines recovery. Am. J. Physiol. 256, C1262–C1266 (1989).

    CAS  PubMed  Google Scholar 

  196. Roberts, P., McGeachie, J. K. & Grounds, M. D. The host environment determines strain-specific differences in the timing of skeletal muscle regeneration: cross-transplantation studies between SJL/J and BALB/c mice. J. Anat. 191, 585–594 (1997).

    PubMed  PubMed Central  Google Scholar 

  197. Rebo, J. et al. A single heterochronic blood exchange reveals rapid inhibition of multiple tissues by old blood. Nat. Commun. 7, 13363 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  198. Sousa-Victor, P., Garcia-Prat, L., Serrano, A. L., Perdiguero, E. & Munoz-Canoves, P. Muscle stem cell aging: regulation and rejuvenation. Trends Endocrinol. Metabol. 26, 287–296 (2015).

    CAS  Google Scholar 

  199. Lukjanenko, L. et al. Aging disrupts muscle stem cell function by impairing matricellular WISP1 secretion from fibro-adipogenic progenitors. Cell Stem Cell 24, 433–446.e7 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  200. Schuler, S. C. et al. Extensive remodeling of the extracellular matrix during aging contributes to age-dependent impairments of muscle stem cell functionality. Cell Rep. 35, 109223 (2021).

    PubMed  Google Scholar 

  201. Kuswanto, W. et al. Poor repair of skeletal muscle in aging mice reflects a defect in local, interleukin-33-dependent accumulation of regulatory T cells. Immunity 44, 355–367 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  202. Wang, Y., Welc, S. S., Wehling-Henricks, M. & Tidball, J. G. Myeloid cell-derived tumor necrosis factor-alpha promotes sarcopenia and regulates muscle cell fusion with aging muscle fibers. Aging Cell 17, e12828 (2018).

    PubMed  PubMed Central  Google Scholar 

  203. Oh, J. et al. Age-associated NF-kappaB signaling in myofibers alters the satellite cell niche and re-strains muscle stem cell function. Aging 8, 2871–2896 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  204. Wang, Y. et al. Aging of the immune system causes reductions in muscle stem cell populations, promotes their shift to a fibrogenic phenotype, and modulates sarcopenia. FASEB J. 33, 1415–1427 (2019).

    CAS  PubMed  Google Scholar 

  205. Tobin, S. W. et al. Delineating the relationship between immune system aging and myogenesis in muscle repair. Aging Cell 20, e13312 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  206. Patsalos, A. et al. In vivo GDF3 administration abrogates aging related muscle regeneration delay following acute sterile injury. Aging Cell 17, e12815 (2018).

    PubMed  PubMed Central  Google Scholar 

  207. Sousa-Victor, P., Neves, J. & Munoz-Canoves, P. Muscle stem cell aging: identifying ways to induce tissue rejuvenation. Mech. Ageing Dev. 188, 111246 (2020).

    CAS  PubMed  Google Scholar 

  208. Zhu, P. et al. The transcription factor Slug represses p16Ink4a and regulates murine muscle stem cell aging. Nat. Commun. 10, 2568 (2019).

    PubMed  PubMed Central  Google Scholar 

  209. Charville, G. W. et al. Ex vivo expansion and in vivo self-renewal of human muscle stem cells. Stem Cell Rep. 5, 621–632 (2015).

    CAS  Google Scholar 

  210. Judson, R. N. et al. Inhibition of methyltransferase Setd7 allows the in vitro expansion of myogenic stem cells with improved therapeutic potential. Cell Stem Cell 22, 177–190.e7 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  211. Sarkar, T. J. et al. Transient non-integrative expression of nuclear reprogramming factors promotes multifaceted amelioration of aging in human cells. Nat. Commun. 11, 1545 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  212. Quarta, M. et al. An artificial niche preserves the quiescence of muscle stem cells and enhances their therapeutic efficacy. Nat. Biotechnol. 34, 752–759 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  213. Brack, A. S. & Rando, T. A. Intrinsic changes and extrinsic influences of myogenic stem cell function during aging. Stem Cell Rev. 3, 226–237 (2007).

    CAS  PubMed  Google Scholar 

  214. Ocampo, A. et al. In vivo amelioration of age-associated hallmarks by partial reprogramming. Cell 167, 1719–1733.e12 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  215. Wang, C. et al. In vivo partial reprogramming of myofibers promotes muscle regeneration by remodeling the stem cell niche. Nat. Commun. 12, 3094 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  216. Sicari, B. M. et al. An acellular biologic scaffold promotes skeletal muscle formation in mice and humans with volumetric muscle loss. Sci. Transl Med. 6, 234ra258 (2014).

    Google Scholar 

  217. Sadtler, K. et al. Developing a pro-regenerative biomaterial scaffold microenvironment requires T helper 2 cells. Science 352, 366–370 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  218. Bencze, M. et al. Proinflammatory macrophages enhance the regenerative capacity of human myoblasts by modifying their kinetics of proliferation and differentiation. Mol. Ther. 20, 2168–2179 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  219. Ho, A. T. V. et al. Prostaglandin E2 is essential for efficacious skeletal muscle stem-cell function, augmenting regeneration and strength. Proc. Natl Acad. Sci. USA 114, 6675–6684 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  220. Riederer, I. et al. Slowing down differentiation of engrafted human myoblasts into immunodeficient mice correlates with increased proliferation and migration. Mol.Ther. 20, 146–154 (2012).

    CAS  PubMed  Google Scholar 

  221. Lafreniere, J. F., Mills, P., Tremblay, J. P. & El Fahime, E. Growth factors improve the in vivo migration of human skeletal myoblasts by modulating their endogenous proteolytic activity. Transplantation 77, 1741–1747 (2004).

    CAS  PubMed  Google Scholar 

  222. Lafreniere, J. F. et al. Growth factor coinjection improves the migration potential of monkey myogenic precursors without affecting cell transplantation success. Cell Transplant. 18, 719–730 (2009).

    PubMed  Google Scholar 

  223. Johnson, S. C., Rabinovitch, P. S. & Kaeberlein, M. mTOR is a key modulator of ageing and age-related disease. Nature 493, 338–345 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  224. Harrison, D. E. et al. Rapamycin fed late in life extends lifespan in genetically heterogeneous mice. Nature 460, 392–395 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  225. Miller, R. A. et al. Rapamycin-mediated lifespan increase in mice is dose and sex dependent and metabolically distinct from dietary restriction. Aging Cell 13, 468–477 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  226. Zhang, Y. et al. Rapamycin extends life and health in C57BL/6 mice. J. Gerontol. A 69, 119–130 (2014).

    CAS  Google Scholar 

  227. Mannick, J. B. et al. mTOR inhibition improves immune function in the elderly. Sci. Transl Med. 6, 268ra179 (2014).

    PubMed  Google Scholar 

  228. Joseph, G. A. et al. Partial inhibition of mTORC1 in aged rats counteracts the decline in muscle mass and reverses molecular signaling associated with sarcopenia. Mol. Cell Biol. 39, e00141-19 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  229. Shavlakadze, T. et al. Short-term low-dose mTORC1 inhibition in aged rats counter-regulates age-related gene changes and blocks age-related kidney pathology. J. Gerontol. A 73, 845–852 (2018).

    CAS  Google Scholar 

  230. Brockhoff, M. et al. Targeting deregulated AMPK/mTORC1 pathways improves muscle function in myotonic dystrophy type I. J. Clin. Invest. 127, 549–563 (2017).

    PubMed  PubMed Central  Google Scholar 

  231. Ramos, F. J. et al. Rapamycin reverses elevated mTORC1 signaling in lamin A/C-deficient mice, rescues cardiac and skeletal muscle function, and extends survival. Sci. Transl Med. 4, 144ra103 (2012).

    PubMed  PubMed Central  Google Scholar 

  232. Bentzinger, C. F. et al. Skeletal muscle-specific ablation of raptor, but not of Rictor, causes metabolic changes and results in muscle dystrophy. Cell Metab. 8, 411–424 (2008).

    CAS  PubMed  Google Scholar 

  233. Risson, V. et al. Muscle inactivation of mTOR causes metabolic and dystrophin defects leading to severe myopathy. J. Cell Biol. 187, 859–874 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  234. Ham, D. J. et al. The neuromuscular junction is a focal point of mTORC1 signaling in sarcopenia. Nat. Commun. 11, 4510 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  235. Segales, J. et al. Sestrin prevents atrophy of disused and aging muscles by integrating anabolic and catabolic signals. Nat. Commun. 11, 189 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  236. Munoz-Canoves, P., Neves, J. & Sousa-Victor, P. Understanding muscle regenerative decline with aging: new approaches to bring back youthfulness to aged stem cells. FEBS J. 287, 406–416 (2020).

    CAS  PubMed  Google Scholar 

  237. Joanisse, S. et al. Exercise conditioning in old mice improves skeletal muscle regeneration. FASEB J. 30, 3256–3268 (2016).

    CAS  PubMed  Google Scholar 

  238. Vinel, C. et al. The exerkine apelin reverses age-associated sarcopenia. Nat. Med. 24, 1360–1371 (2018).

    CAS  PubMed  Google Scholar 

  239. Wang, J. et al. Engineered skeletal muscles for disease modeling and drug discovery. Biomaterials 221, 119416 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  240. Lancaster, M. A. & Knoblich, J. A. Generation of cerebral organoids from human pluripotent stem cells. Nat. Protoc. 9, 2329–2340 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  241. Carnes, M. E. & Pins, G. D. Skeletal muscle tissue engineering: biomaterials-based strategies for the treatment of volumetric muscle loss. Bioengineering 7, 85 (2020).

    CAS  PubMed Central  Google Scholar 

  242. Alsharidah, M. et al. Primary human muscle precursor cells obtained from young and old donors produce similar proliferative, differentiation and senescent profiles in culture. Aging Cell 12, 333–344 (2013).

    CAS  PubMed  Google Scholar 

  243. Bigot, A. et al. Replicative aging down-regulates the myogenic regulatory factors in human myoblasts. Biol. Cell 100, 189–199 (2008).

    CAS  PubMed  Google Scholar 

  244. Nehlin, J. O., Just, M., Rustan, A. C. & Gaster, M. Human myotubes from myoblast cultures undergoing senescence exhibit defects in glucose and lipid metabolism. Biogerontology 12, 349–365 (2011).

    CAS  PubMed  Google Scholar 

  245. Del Carmen Ortuno-Costela, M., Garcia-Lopez, M., Cerrada, V. & Gallardo, M. E. iPSCs: a powerful tool for skeletal muscle tissue engineering. J. Cell Mol. Med. 23, 3784–3794 (2019).

    PubMed  PubMed Central  Google Scholar 

  246. Thomson, J. A. et al. Embryonic stem cell lines derived from human blastocysts. Science 282, 1145–1147 (1998).

    CAS  PubMed  Google Scholar 

  247. Takahashi, K. et al. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell 131, 861–872 (2007).

    CAS  PubMed  Google Scholar 

  248. Lagha, M. et al. Pax3 regulation of FGF signaling affects the progression of embryonic progenitor cells into the myogenic program. Genes Dev. 22, 1828–1837 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  249. Barberi, T. et al. Derivation of engraftable skeletal myoblasts from human embryonic stem cells. Nat. Med. 13, 642–648 (2007).

    CAS  PubMed  Google Scholar 

  250. Xu, C. et al. A zebrafish embryo culture system defines factors that promote vertebrate myogenesis across species. Cell 155, 909–921 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  251. Borchin, B., Chen, J. & Barberi, T. Derivation and FACS-mediated purification of PAX3+/PAX7+ skeletal muscle precursors from human pluripotent stem cells. Stem Cell Rep. 1, 620–631 (2013).

    CAS  Google Scholar 

  252. Hwang, Y. et al. Directed in vitro myogenesis of human embryonic stem cells and their in vivo engraftment. PLoS ONE 8, e72023 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  253. Engler, A. J., Sen, S., Sweeney, H. L. & Discher, D. E. Matrix elasticity directs stem cell lineage specification. Cell 126, 677–689 (2006).

    CAS  PubMed  Google Scholar 

  254. Maffioletti, S. M. et al. Efficient derivation and inducible differentiation of expandable skeletal myogenic cells from human ES and patient-specific iPS cells. Nat. Protoc. 10, 941–958 (2015).

    CAS  PubMed  Google Scholar 

  255. Darabi, R. et al. Human ES- and iPS-derived myogenic progenitors restore DYSTROPHIN and improve contractility upon transplantation in dystrophic mice. Cell Stem Cell 10, 610–619 (2012). This study shows that conditional expression of PAX7 in human ES cells/iPS cells produces large quantities of myogenic precursors, which engraft efficiently.

    CAS  PubMed  PubMed Central  Google Scholar 

  256. Roca, I., Requena, J., Edel, M. J. & Alvarez-Palomo, A. B. Myogenic precursors from iPS cells for skeletal muscle cell replacement therapy. J. Clin. Med. 4, 243–259 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  257. Mondragon-Gonzalez, R. & Perlingeiro, R. C. R. Recapitulating muscle disease phenotypes with myotonic dystrophy 1 induced pluripotent stem cells: a tool for disease modeling and drug discovery. Dis. Model. Mech. https://doi.org/10.1242/dmm.034728 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  258. van der Wal, E. et al. Large-scale expansion of human iPSC-derived skeletal muscle cells for disease modeling and cell-based therapeutic strategies. Stem Cell Rep. 10, 1975–1990 (2018).

    Google Scholar 

  259. Kim, H. et al. Genomic safe harbor expression of PAX7 for the generation of engraftable myogenic progenitors. Stem Cell Rep. 16, 10–19 (2021).

    CAS  Google Scholar 

  260. Maffioletti, S. M. et al. Three-dimensional human iPSC-derived artificial skeletal muscles model muscular dystrophies and enable multilineage tissue engineering. Cell Rep. 23, 899–908 (2018). This work reports the generation of 3D artificial skeletal muscle tissue (a muscle organoid) from human PS cells from patients with Duchenne, limb-girdle and congenital muscular dystrophies.

    CAS  PubMed  PubMed Central  Google Scholar 

  261. Selvaraj, S. et al. Screening identifies small molecules that enhance the maturation of human pluripotent stem cell-derived myotubes. eLife 8, e47970 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  262. Aloysius, A., DasGupta, R. & Dhawan, J. The transcription factor Lef1 switches partners from β-catenin to Smad3 during muscle stem cell quiescence. Sci. Signal. 11, eaan3000 (2018).

    PubMed  Google Scholar 

  263. Bjornson, C. R. et al. Notch signaling is necessary to maintain quiescence in adult muscle stem cells. Stem Cell 30, 232–242 (2012).

    CAS  Google Scholar 

  264. Mourikis, P. et al. A critical requirement for notch signaling in maintenance of the quiescent skeletal muscle stem cell state. Stem Cell 30, 243–252 (2012).

    CAS  Google Scholar 

  265. Lala-Tabbert, N., AlSudais, H., Marchildon, F., Fu, D. & Wiper-Bergeron, N. CCAAT/enhancer-binding protein beta promotes muscle stem cell quiescence through regulation of quiescence-associated genes. Stem Cell (2020).

  266. Xie, L. et al. Transient HIF2A inhibition promotes satellite cell proliferation and muscle regeneration. J. Clin. Invest. 128, 2339–2355 (2018).

    PubMed  PubMed Central  Google Scholar 

  267. Hosoyama, T., Nishijo, K., Prajapati, S. I., Li, G. & Keller, C. Rb1 gene inactivation expands satellite cell and postnatal myoblast pools. J. Biol. Chem. 286, 19556–19564 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  268. Sato, T., Yamamoto, T. & Sehara-Fujisawa, A. miR-195/497 induce postnatal quiescence of skeletal muscle stem cells. Nat. Commun. 5, 4597 (2014).

    CAS  PubMed  Google Scholar 

  269. Baghdadi, M. B. et al. Notch-induced miR-708 antagonizes satellite cell migration and maintains quiescence. Cell Stem Cell 23, 859–868.e5 (2018).

    CAS  PubMed  Google Scholar 

  270. Chenette, D. M. et al. Targeted mRNA decay by RNA binding protein AUF1 regulates adult muscle stem cell fate, promoting skeletal muscle integrity. Cell Rep. 16, 1379–1390 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  271. Abou-Khalil, R. & Brack, A. S. Muscle stem cells and reversible quiescence: the role of sprouty. Cell Cycle 9, 2575–2580 (2010).

    CAS  PubMed  Google Scholar 

  272. Mizuno, S. et al. Inhibition of ADAM10 in satellite cells accelerates muscle regeneration following muscle injury. J. Orthop. Res. 36, 2259–2265 (2018).

    CAS  Google Scholar 

  273. McCroskery, S., Thomas, M., Maxwell, L., Sharma, M. & Kambadur, R. Myostatin negatively regulates satellite cell activation and self-renewal. J. Cell Biol. 162, 1135–1147 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  274. Zhou, S. et al. Paxbp1 controls a key checkpoint for cell growth and survival during early activation of quiescent muscle satellite cells. Proc. Natl Acad. Sci. USA 118, e2021093118 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  275. Elabd, C. et al. Oxytocin is an age-specific circulating hormone that is necessary for muscle maintenance and regeneration. Nat. Commun. 5, 4082 (2014).

    CAS  PubMed  Google Scholar 

  276. Chang, J. et al. Clearance of senescent cells by ABT263 rejuvenates aged hematopoietic stem cells in mice. Nat. Med. 22, 78–83 (2016).

    CAS  PubMed  Google Scholar 

  277. Gibala, M. J., Little, J. P., Macdonald, M. J. & Hawley, J. A. Physiological adaptations to low-volume, high-intensity interval training in health and disease. J. Physiol. 590, 1077–1084 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  278. O’Reilly, C., McKay, B., Phillips, S., Tarnopolsky, M. & Parise, G. Hepatocyte growth factor (HGF) and the satellite cell response following muscle lengthening contractions in humans. Muscle Nerve 38, 1434–1442 (2008).

    PubMed  Google Scholar 

  279. Crameri, R. M. et al. Changes in satellite cells in human skeletal muscle after a single bout of high intensity exercise. J. Physiol. 558, 333–340 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  280. Mackey, A. L. et al. Assessment of satellite cell number and activity status in human skeletal muscle biopsies. Muscle Nerve 40, 455–465 (2009).

    PubMed  Google Scholar 

  281. Ingjer, F. Effects of endurance training on muscle fibre ATP-ase activity, capillary supply and mitochondrial content in man. J. Physiol. 294, 419–432 (1979).

    CAS  PubMed  PubMed Central  Google Scholar 

  282. Bazgir, B., Fathi, R., Rezazadeh Valojerdi, M., Mozdziak, P. & Asgari, A. Satellite cells contribution to exercise mediated muscle hypertrophy and repair. Cell J. 18, 473–484 (2017).

    PubMed  Google Scholar 

  283. Kadi, F. et al. The effects of heavy resistance training and detraining on satellite cells in human skeletal muscles. J. Physiol. 558, 1005–1012 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  284. Garg, K. & Boppart, M. D. Influence of exercise and aging on extracellular matrix composition in the skeletal muscle stem cell niche. J. Appl. Physiol. 121, 1053–1058 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  285. Mackey, A. L. et al. Sequenced response of extracellular matrix deadhesion and fibrotic regulators after muscle damage is involved in protection against future injury in human skeletal muscle. FASEB J. 25, 1943–1959 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  286. Smith, H. K. & Merry, T. L. Voluntary resistance wheel exercise during post-natal growth in rats enhances skeletal muscle satellite cell and myonuclear content at adulthood. Acta Physiol. 204, 393–402 (2012).

    CAS  Google Scholar 

  287. Song, W., Kwak, H. B. & Lawler, J. M. Exercise training attenuates age-induced changes in apoptotic signaling in rat skeletal muscle. Antioxid. Redox Signal. 8, 517–528 (2006).

    CAS  PubMed  Google Scholar 

  288. Melov, S., Tarnopolsky, M. A., Beckman, K., Felkey, K. & Hubbard, A. Resistance exercise reverses aging in human skeletal muscle. PLoS ONE 2, e465 (2007).

    PubMed  PubMed Central  Google Scholar 

  289. Lacraz, G. et al. Increased stiffness in aged skeletal muscle impairs muscle progenitor cell proliferative activity. PLoS ONE 10, e0136217 (2015).

    PubMed  PubMed Central  Google Scholar 

  290. Carroll, C. C. et al. The effect of chronic treadmill exercise and acetaminophen on collagen and cross-linking in rat skeletal muscle and heart. Am. J. Physiol. Regul. Integr. Comp. Physiol. 308, R294–R299 (2015).

    CAS  PubMed  Google Scholar 

  291. Gosselin, L. E., Adams, C., Cotter, T. A., McCormick, R. J. & Thomas, D. P. Effect of exercise training on passive stiffness in locomotor skeletal muscle: role of extracellular matrix. J. Appl. Physiol. 85, 1011–1016 (1998).

    CAS  PubMed  Google Scholar 

  292. Marcus, R. L., Addison, O., Kidde, J. P., Dibble, L. E. & Lastayo, P. C. Skeletal muscle fat infiltration: impact of age, inactivity, and exercise. J. Nutr. Health Aging 14, 362–366 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  293. Zanandrea, V. et al. Interventions against sarcopenia in older persons. Curr. Pharm. Des. 20, 5983–6006 (2014).

    PubMed  Google Scholar 

  294. Shefer, G., Rauner, G., Stuelsatz, P., Benayahu, D. & Yablonka-Reuveni, Z. Moderate-intensity treadmill running promotes expansion of the satellite cell pool in young and old mice. FEBS J. 280, 4063–4073 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors thank all reviewers for their insightful evaluation of and feedback for this Review. They also thank J. Neves for helpful comments on the manuscript. Furthermore, they thank many of their colleagues in the muscle regeneration field whose work inspired them for this Review. They regret that due to reference constraints, they could not cite all of the worthy articles on this subject. Work in the authors’ laboratory was supported in part by the Spanish MINECO (RTI2018-096068), ERC-AdG-741966, LaCaixa-HEALTH-HR17-00040, MDA, UPGRADE-H2020-825825, MWRF, FundacióLaMaratóTV3, AFM and DPP-Spain (P.M.-C). P.M.-C’s laboratories at Pompeu Fabra University and the Spanish National Center for Cardiovascular Research are recipients of grants from the María de Maeztu Units of Excellence programme to Pompeu Fabra University (MDM-2014-0370) and the Severo Ochoa Centers of Excellence programme to the Spanish National Center for Cardiovascular Research (SEV-2015-0505) for P.M.C. Work in the P.S.-V. laboratory was supported by the Fundação para a Ciência e Tecnologia Project (PTDC/MED-OUT/8010/2020), an EMBO installation grant (IG4448) and the ‘la Caixa’ Foundation for a junior leader fellowship (LCF/BQ/PI19/11690006) to P.S.-V. L.G.-P. was supported by an EMBO long-term fellowship (ALTF 420-2017), a Benjamin Pearl fellowship, and a CIHR fellowship (201910MFE-430959-284655).

Author information

Authors and Affiliations

Authors

Contributions

The authors contributed equally to all aspects of the article.

Corresponding authors

Correspondence to Pedro Sousa-Victor, Laura García-Prat or Pura Muñoz-Cánoves.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information

Nature Reviews Molecular Cell Biology thanks the anonymous reviewers for their contribution to the peer review of this work.

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary information

Glossary

Basal lamina

The inner layer of the basement membrane (composed of extracellular matrix proteins) that is adjacent to the muscle sarcolemma.

Hepatocyte growth factor (HGF) activator

(HGFA). A serine protease that activates HGF by converting it from a single chain to a heterodimeric form.

Duchenne muscular dystrophy

A genetic disorder characterized by progressive muscle degeneration and weakness due to a mutation in a gene coding for the protein dystrophin.

Myofibre type

The type of myosin expressed by each muscle fibre (myofibre) determines the myofibre type, which broadly categorizes them into types 1 and 2 myofibres, which in turn differ in the type of metabolism used for energy. Myofibres can be further subdivided into four types, identified by the expression of four myosin heavy chain (MyHC) isotypes: MyHC-2b, MyHC-2x, MyHC-2a and MyHC-1. Myofibres are also broadly classified as ‘slow-twitch‘ (type 1 MyHC) and ‘fast-twitch’ (type 2 MyHC) myofibres and differ in energy production: types 1 and 2a fibres primarily use oxidative metabolism, and types 2x and 2b myofibres primarily rely on glycolytic metabolism. These fibre types are dynamically interchangeable as an adaptation to alterations in muscle function and metabolism.

Asymmetric cell division

A process whereby the asymmetric inheritance of cellular components (for example, proteins, RNAs) during mitosis defines distinct fates for each daughter cell. This division mode is used by stem and progenitor cells in different tissues. Satellite cells in homeostasis usually undergo asymmetric cell division, which generates one self-renewing stem cell and one differentiating cell. By contrast, symmetric cell division occurs once satellite cells are activated and generates one type of cell (either two self-renewing stem cells or two differentiating cells).

CIP/KIP family of cell cycle inhibitors

Cyclin-dependent kinase (CDK)-interacting protein/kinase inhibitory protein family of cyclin-dependent kinase inhibitors, which bind both cyclin and CDK through a conserved amino-terminal domain. The family is composed of p21CIP1/WAF1, p27KIP1 and p57KIP2.

Schwann cells

Glial cells of the peripheral nervous system that surround axons of motor and sensory neurons and produce a myelin sheath.

Smooth muscle–mesenchymal cells

(SMMCs). A subpopulation of ITGA7+ cells identified in the skeletal muscle that are distinct from satellite cells and express markers of the mesenchymal and smooth muscle cell lineages.

FOXO

A subgroup of the forkhead family of transcription factors that orchestrate programmes of gene expression that regulate crucial cellular processes, including cell homeostasis, cell cycle progression, oxidative stress responses and metabolism and that share the characteristic of being regulated by the insulin–PI3K–AKT signalling pathway.

Calcitonin receptor

(CALCR). Seven-membrane-spanning domain G protein-coupled receptor that binds the peptide hormone calcitonin. It is involved in the maintenance of calcium homeostasis.

Tenocytes

Elongated fibroblast type tendon cells responsible for synthesis and turnover of tendon fibres and production of extracellular matrix components.

Syndecans

Single transmembrane domain proteoglycans with multiple heparan sulfate and chondroitin sulfate chains capable of interacting with multiple ligands and acting as co-receptors for G-protein-coupled receptors.

AMP-activated protein kinase

(AMPK). 5′-Adenosine monophosphate-activated protein kinase, a regulator of cellular energy homeostasis.

MYC

Transcription factor of the basic helix–loop–helix family encoded by the proto-oncogene MYC. It is constitutively expressed in several cancers and is associated with the promotion of cell proliferation.

E2F

Transcription factors, binding to the TTTCCCGC consensus sequence in target promoters, involved in the regulation of the cell cycle and DNA synthesis.

Dicer

An RNase III endonuclease that processes microRNA precursors into functional 21–23-nucleotode RNAs that are subsequently incorporated into the RNA-induced silencing complex.

Tristetraprolin

(TTP). An RNA-binding protein that binds the 3ʹ untranslated region of target RNAs, leading to their rapid decay through the recruitment of cytoplasmic RNA-degradation machinery.

mRNA decay

A step in the process of gene expression (from gene to protein) consisting of the degradation or destruction of mRNA.

p38 mitogen-activated protein kinases

(MAPKs). A class of MAPKs (specific to the amino acids serine and threonine) that are responsive to stress stimuli. There are four isoforms of the p38 family of MAPKs (α, β, γ and δ), which have a key function in transducing extracellular signals from outside the cell into the nucleus, through a series of phosphorylation events, ultimately activating effectors of gene expression.

YAP

Yes-associated protein, a transcription factor that activates genes involved in cell proliferation and suppresses apoptotic genes. It also regulates genes in response to mechanotransduction. YAP is inhibited in the Hippo signalling pathway.

NAD+

Nicotinamide adenine dinucleotide (NAD), a coenzyme that is a hydrogen carrier central to metabolism, exists in two forms: an oxidized form (NAD+) and a reduced form (NADH). The main role of NAD+ in metabolism is the transfer of electrons from one molecule to another in redox reactions.

mTORC1

Mechanistic target of rapamycin complex 1, part of a protein complex that regulates protein synthesis in response to nutrients and metabolic products in the cell.

mTORC2

Mechanistic target of rapamycin complex 2, which is involved in the regulation of cell proliferation, survival and migration.

Unfolded protein response

A cellular stress response activated by the accumulation of unfolded or misfolded proteins in the endoplasmic reticulum, aiming to restore normal function.

Dystrophin glycoprotein complex

A large multiprotein complex of the skeletal and cardiac muscle membrane, with both mechanically stabilizing and signalling roles in mediating interactions between the cytoskeleton, membrane and extracellular matrix. Mutations in dystrophin or additional dystrophin glycoprotein complex genes, the sarcoglycans, lead to muscular dystrophies.

Negative elongation factor

Protein complex that blocks transcription by pausing RNA polymerase II activity, preventing early transcript elongation.

Necroptosis

A programmed form of necrosis, or inflammatory cell death, usually associated with cellular damage or infiltration by pathogens. Receptor-interacting protein kinase 3 (RIP3K) is essential for necroptosis.

Gnai2

The gene encoding GTP-binding regulatory protein Gi α2 chain. It regulates proliferation and differentiation of several cell types, including satellite cells.

Neuromuscular junctions

(NMJs). The synapses between motor neurons and muscle fibres.

Regulatory T cells

(Treg cells). A subpopulation of T cells that modulate the function of other immune cells. They are identified by the expression of CD4, FOXP3 and CD25.

Senescence-associated secretory phenotype

Collection of secreted molecules derived from senescent cells that includes inflammatory cytokines, immune modulators, growth factors and proteases with paracrine action on surrounding cells.

Inflammageing

A state of low-grade chronic inflammation that develops in aged organisms.

Heterochronic parabiosis

An experimental procedure that combines two living organisms which are joined together surgically to develop a single, shared physiological system. In the context of ageing, when two animals of different ages are joined to test for systemic regulators of aspects of ageing or age-related diseases.

Yamanaka factors

OCT3/4, SOX2, KLF4 and MYC (OSKM factors) are a group of transcription factors identified by Yamanaka as capable of creating induced pluripotent stem cells.

Resistance training

A type of exercise where the muscles contract against an external resistance, promoting strength and hypertrophy.

Endurance training

A type of exercise that promotes adaptation of the skeletal muscle to aerobic metabolism, reflected by an increased oxidative capacity.

Genomic safe harbour

A genomic location where the integration of new genetic elements does not cause alterations of the host genome that pose risk to the cell, allowing the predictable function of the inserted transgenes.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sousa-Victor, P., García-Prat, L. & Muñoz-Cánoves, P. Control of satellite cell function in muscle regeneration and its disruption in ageing. Nat Rev Mol Cell Biol 23, 204–226 (2022). https://doi.org/10.1038/s41580-021-00421-2

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41580-021-00421-2

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing